You are on page 1of 262

Title Mixing and deposition of sediment-laden buoyant jets

Advisor(s) Lam, KM

Author(s) Chan, Shu-ning.; .

Chan, S. []. (2013). Mixing and deposition of sediment-


laden buoyant jets. (Thesis). University of Hong Kong, Pokfulam,
Citation Hong Kong SAR. Retrieved from
http://dx.doi.org/10.5353/th_b5060572

Issued Date 2013

URL http://hdl.handle.net/10722/188746

The author retains all proprietary rights, (such as patent rights)


Rights and the right to use in future works.
Abstract of thesis entitled

Mixing and Deposition of Sediment-Laden Buoyant Jets

submitted by

CHAN Shu Ning

for the degree of Doctor of Philosophy


at the University of Hong Kong
in January 2013

Sediment-laden turbulent buoyant jets are commonly encountered in the nat-


ural and man-made environments. Examples of sediment-laden buoyant jets
include volcanic eruptions, deep ocean hydrothermal vents (black smokers),
ocean dumping of dredged spoils and sludge, and submarine discharge of wastew-
ater effluent. It is important to understand the fluid mechanics of sediment jets
for environmental impact assessment, and yet there is currently no general model
for predicting the mixing of sediment-laden jets. This study reports a theoretical
and experimental investigation the sediment mixing, fall-out and deposition from
sediment-laden buoyant jets.
It is well known that turbulence generates fluctuations to the particle mo-
tion, modulating the particle settling velocity. A general three-dimensional (3D)
stochastic particle tracking model is developed to predict the particle settling
out and deposition from a sediment-laden jet. Particle velocity fluctuations are
modelled by a Lagrangian velocity autocorrelation function that accounts for
the loitering and trapping of sediment particles in turbulent eddies which results
in the reduction of settling velocity. The model is validated against results of
independent experimental studies. Consistent with basic experiments using grid-
generated turbulence, the model predicts that the apparent settling velocity can
be reduced by as much as 30% of the stillwater settling velocity.
The mixing and deposition of sediment-laden horizontal momentum jets are
studied using laboratory experiments and 3D computational fluid dynamics (CFD)
modelling. It is shown that there is a significant settling velocity reduction up
to about 25-35%, dependent on jet turbulent fluctuations and particle proper-
ties. The CFD approach necessitates an ad hoc adjustment/reduction on settling
velocity and lacks generality. Using classical solutions of mean velocity, and tur-
bulent fluctuation and dissipation rate profiles derived from CFD solutions, 3D
particle tracking model predictions of sediment deposition and concentration pro-
files are in excellent agreement with measured data over a wide range of jet flow
and particle properties. Unlike CFD calculations, the present method does not
require any a priori adjustment of particle settling velocity.
A general particle tracking model for predicting sediment fall-out and deposi-
tion from an arbitrarily inclined buoyant jets in stagnant ambient is successfully
developed. The model incorporates the three flow regimes affecting the sediment
dynamics in a buoyant jet, namely turbulent jet flow, jet entrainment-induced
external flow and surface spreading current. The jet mean flow velocity is de-
termined using a well-validated jet integral model. The external jet-induced
irrotational flow field is computed by a distribution of point sinks along the jet
trajectory. The surface spreading current is predicted using an integral model
accounting for the interfacial shear. The model is validated against experimental
data of sediment deposition from vertical and horizontal sediment-laden buoyant
jets.
Mixing and Deposition of
Sediment-Laden Buoyant Jets

CHAN Shu Ning

A thesis submitted
in partial fulfillment of the requirements
for the degree of Doctor of Philosophy
at The University of Hong Kong

January 2013
2
Declaration

I declare that this thesis represents my own work, except where due acknowl-
edgment is made, and that it has not been previously included in a thesis, dis-
sertation or report submitted to this University or to any other institution for a
degree, diploma or other qualifications.

...............................................
(CHAN Shu Ning)

i
ii DECLARATION
Acknowledgement

I would like to express my sincere gratitude to my supervisor, Prof. Joseph H.W.


Lee, for his guidance and encouragement throughout my study. Throughout the
years, I have learned a lot from his rigorous attitude towards research and his
commitment on engineering a better world. His critical comments and insightful
suggestions are always useful to my research and future work. I am the most
grateful that he spared his busy times with extreme patience in commenting and
editing my thesis and papers.
The advice and suggestions from Dr. K.M. Lam on my study are beneficial
and gratefully acknowledged.
I express my gratitude to all my colleagues in the Croucher Laboratory of
Environmental Hydraulics, The University of Hong Kong. Special thank goes to
Dr. Ken Lee, who shared with me his experimental expertise, without which this
work would not have been completed successfully. The professional and efficient
technical support from the lab technicians, Mr. C.H. Tong and Mr. T.O. Chan,
is the most appreciated. Stimulating discussions with Dr. P. Liu and Mr. Chris
Lai on experimental techniques are always gainful. Dr. Adrian Lai provided
me fruitful ideas on theoretical aspects of the subject. The work on Project
WATERMAN with Dr. David Choi, Dr. Ken Wong, Dr. Wai Thoe and other team
members has broadened my horizon on the challenging environmental problems.
The days and nights in the Laboratory would be memorable in my life.
I would like to thank Prof. X.Y. Li and Mr. Keith Wong for allowing me to
use their laboratory facilities. My gratitude also goes to Ms. Candice Fong in
the Department of Civil Engineering, HKU for her administrative support.
The financial support from the Area of Excellence (AoE) in Marine Environ-
mental Research and Innovative Technology (MERIT), Hong Kong University
Grants Committee on my study is gratefully acknowledged.
Last but not least, I would like to thank my family for their wholehearted
support during my study journey.

iii
iv
Contents

Declaration i

Acknowledgement iii

List of Symbols xxiii

1 Introduction 1
1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1.1 Volcanic eruption . . . . . . . . . . . . . . . . . . . . . . . 1
1.1.2 Submarine hydrothermal vents - black smokers . . . . . 2
1.1.3 Marine disposal of dredged spoils and sludge . . . . . . . . 4
1.1.4 Wastewater disposal in the marine environment . . . . . . 5
1.2 Motivation and objective of this study . . . . . . . . . . . . . . . 6
1.3 Outline of the present work . . . . . . . . . . . . . . . . . . . . . 9

2 Literature Review 11
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.2 Single-phase buoyant jet in stagnant fluid . . . . . . . . . . . . . . 11
2.2.1 The mean flow properties . . . . . . . . . . . . . . . . . . 11
2.2.2 Jet/plume turbulence . . . . . . . . . . . . . . . . . . . . . 13
2.2.3 Jet induced external flow . . . . . . . . . . . . . . . . . . . 13
2.2.4 Gravitational spreading current induced by jet impingement 14
2.3 Sediment-laden jets . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.3.1 Vertical downward sediment-laden jets . . . . . . . . . . . 14
2.3.2 Vertical upward sediment-laden jets . . . . . . . . . . . . . 15
2.3.3 Horizontal sediment-laden buoyant jets . . . . . . . . . . . 16
2.3.4 Sediment-laden jet applications in ocean dumping . . . . . 17
2.3.5 Prediction of sediment impact of ocean outfalls . . . . . . 18
2.3.6 Summary of previous studies of sediment-laden jets . . . . 18

v
2.4 The equation of motion of sediment particle . . . . . . . . . . . . 19
2.4.1 Sediment settling velocity . . . . . . . . . . . . . . . . . . 19
2.4.2 Basset history force . . . . . . . . . . . . . . . . . . . . . . 19
2.5 Particle settling in turbulent flows . . . . . . . . . . . . . . . . . . 20
2.6 Mathematical modelling of two-phase flows . . . . . . . . . . . . . 22
2.6.1 Eulerian modelling . . . . . . . . . . . . . . . . . . . . . . 22
2.6.2 Lagrangian modelling . . . . . . . . . . . . . . . . . . . . . 23
2.7 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

3 Theory: Particle Tracking Model 27


3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.2 The governing equation of particle motion . . . . . . . . . . . . . 27
3.2.1 Components of the equation of motion . . . . . . . . . . . 27
3.2.2 Solution of the equation of motion . . . . . . . . . . . . . 31
3.3 The autocorrelation function of a sediment particle in turbulence . 31
3.3.1 Taylors autocorrelation function . . . . . . . . . . . . . . 31
3.3.2 Derivation of the equivalent Eulerian particle velocity au-
tocorrelation function . . . . . . . . . . . . . . . . . . . . . 32
3.3.3 The autocorrelation of sediment particle in turbulence . . 34
3.3.4 Generation of turbulent velocity fluctuations . . . . . . . . 38
3.3.5 Physical interpretation of the autocorrelation function . . . 38
3.4 Model validation . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.4.1 Accelerated motion of a particle in stagnant fluid . . . . . 42
3.4.2 Motion of particles in homogeneously vertically oscillating
fluid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.4.3 Motion of particles in homogeneous grid turbulence . . . . 49
3.4.4 Vertically downward sediment-laden jets . . . . . . . . . . 56
3.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

4 Experiments 65
4.1 Experimental set-up . . . . . . . . . . . . . . . . . . . . . . . . . 65
4.1.1 Experimental tank . . . . . . . . . . . . . . . . . . . . . . 65
4.1.2 Jet discharge system . . . . . . . . . . . . . . . . . . . . . 65
4.1.3 Sediment supply system . . . . . . . . . . . . . . . . . . . 67
4.1.4 Bottom sediment collection and measurement . . . . . . . 68
4.2 Particle properties . . . . . . . . . . . . . . . . . . . . . . . . . . 74
4.2.1 Glass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

vi
4.2.2 Polymethylmethacrylate, IP3 . . . . . . . . . . . . . . . . 74
4.2.3 Melamine formaldehyde, MF . . . . . . . . . . . . . . . . . 75
4.2.4 Measurement of settling velocity and determination of par-
ticle equivalent diameter . . . . . . . . . . . . . . . . . . 75
4.3 Flow visualization . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
4.4 Fluid density and viscosity calculation . . . . . . . . . . . . . . . 76
4.5 Summary of experiments . . . . . . . . . . . . . . . . . . . . . . . 77
4.6 Cross-sectional sediment concentration measurement . . . . . . . 80
4.7 Measurement of jet turbulent velocity . . . . . . . . . . . . . . . . 82
4.7.1 Particle Imaging Velocimetry . . . . . . . . . . . . . . . . 82
4.7.2 Mean flow and turbulent fluctuations . . . . . . . . . . . . 83
4.7.3 Experimental support of trapping effect in jet turbulence . 86
4.7.4 Experimental support of negligible cross-correlations be-
tween partial velocity derivatives . . . . . . . . . . . . . . 90
4.8 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92

5 Modelling of Sediment-Laden Horizontal Momentum Jets 93


5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
5.2 Horizontal sediment-laden momentum jet . . . . . . . . . . . . . . 93
5.3 Experimental observation . . . . . . . . . . . . . . . . . . . . . . 94
5.3.1 Flow visualization . . . . . . . . . . . . . . . . . . . . . . . 94
5.3.2 Sediment bottom deposition . . . . . . . . . . . . . . . . . 95
5.4 3D Computational Fluid Dynamics modelling . . . . . . . . . . . 104
5.5 Particle tracking modelling . . . . . . . . . . . . . . . . . . . . . . 109
5.5.1 Mean flow velocity . . . . . . . . . . . . . . . . . . . . . . 109
5.5.2 Turbulence quantities . . . . . . . . . . . . . . . . . . . . . 109
5.5.3 Particle size . . . . . . . . . . . . . . . . . . . . . . . . . . 110
5.5.4 Computational procedures . . . . . . . . . . . . . . . . . . 111
5.6 Particle tracking modelling results . . . . . . . . . . . . . . . . . . 112
5.6.1 1D sediment deposition profiles . . . . . . . . . . . . . . . 112
5.6.2 2D sediment deposition profiles . . . . . . . . . . . . . . . 118
5.6.3 Importance of different forces in the equation of motion to
the prediction of deposition profile . . . . . . . . . . . . . 118
5.6.4 Cross-sectional sediment concentration profiles . . . . . . . 128
5.7 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138

6 A General Model of Sediment-Laden Buoyant Jets 141


6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141

vii
6.2 The flow regimes in buoyant jets . . . . . . . . . . . . . . . . . . . 141
6.2.1 The buoyant jet flow . . . . . . . . . . . . . . . . . . . . . 142
6.2.2 The external flow . . . . . . . . . . . . . . . . . . . . . . . 148
6.2.3 The surface spreading layer . . . . . . . . . . . . . . . . . 152
6.3 Particle tracking model . . . . . . . . . . . . . . . . . . . . . . . . 156
6.4 Vertically upward sediment-laden buoyant jet . . . . . . . . . . . 157
6.4.1 General observation . . . . . . . . . . . . . . . . . . . . . . 157
6.4.2 Sediment fall out from jet margin . . . . . . . . . . . . . . 159
6.4.3 Sediment fall out from spreading current . . . . . . . . . . 165
6.4.4 Comparison of the Full Model and the Simplified Model . 167
6.5 Horizontal sediment-laden buoyant jet . . . . . . . . . . . . . . . 169
6.5.1 Experiments of Li (2006) and Lee (2010) . . . . . . . . . . 169
6.5.2 CFD model prediction . . . . . . . . . . . . . . . . . . . . 172
6.5.3 Particle model prediction . . . . . . . . . . . . . . . . . . . 176
6.5.4 Plastic particle experiments: Lee (2010), Cuthbertson and
Davies (2008) . . . . . . . . . . . . . . . . . . . . . . . . . 185
6.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188

7 Conclusion 189
7.1 Summary of this study . . . . . . . . . . . . . . . . . . . . . . . . 189
7.2 Recommendations for future work . . . . . . . . . . . . . . . . . . 191
7.3 Application examples . . . . . . . . . . . . . . . . . . . . . . . . . 192

A Numerical Solution of the Equation of Particle Motion 209


A.1 General numerical solution . . . . . . . . . . . . . . . . . . . . . . 209
A.2 Numerical implementation of the Basset force term . . . . . . . . 210

B 3D Computation Fluid Dynamics Modelling 215


B.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
B.2 Governing equations . . . . . . . . . . . . . . . . . . . . . . . . . 215
B.3 Turbulence closure . . . . . . . . . . . . . . . . . . . . . . . . . . 216
B.4 Model of horizontal sediment-laden jet . . . . . . . . . . . . . . . 217
B.4.1 Grid configuration . . . . . . . . . . . . . . . . . . . . . . 217
B.4.2 Boundary and initial conditions . . . . . . . . . . . . . . . 218
B.4.3 Computational details . . . . . . . . . . . . . . . . . . . . 219
B.4.4 Model results . . . . . . . . . . . . . . . . . . . . . . . . . 219
B.5 Model for spreading current . . . . . . . . . . . . . . . . . . . . . 224

viii
C Particle Settling Velocity in Ethanol-Water Mixture 227

ix
x
List of Figures

1.1 Eruption of Mayon Volcano, the Philippines in 1984, showing the


volcanic plume of hot gases and volcanic ashes, and streams of
pyroclastic flows running down the slope. . . . . . . . . . . . . . . 3
1.2 A deep ocean hydrothermal vent - black smoker in Atlantic
Ocean. Precipitated minerals are deposited around the vent. . . . 3
1.3 Dumping of dredged spoils from a barge. . . . . . . . . . . . . . . 4
1.4 Sewage outfalls in Hong Kong (preliminary, primary and sec-
ondary treatment), marine dumping sites and sensitive receivers
(bathing beaches, fish culture zones and marine reserves). The
solid line is the transfer tunnels of HATS Stage 1 in operation.
The dashed line is the transfer tunnels of HATS Stage 2A com-
mencing in 2014. . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.5 3D depiction of the diffuser of Harbour Area Treatment Scheme
(HATS) outfall (1.2km long, 24 rosette risers, each with 8 jets).
Two of the rosette risers are shown in detail. . . . . . . . . . . . . 7
1.6 Experimental image of a sediment-laden jet in this study. Jet
velocity = 0.79m/s; glass particles of 180 m and settling velocity
2.0 cm/s are used. . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2.1 A single phase buoyant jet. . . . . . . . . . . . . . . . . . . . . . . 12


2.2 Trapping of sediment particle in a forced vortex adapted from
Nielsen (1984). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

3.1 Drag coefficient-particle Reynolds number relationship by Clift


and Gauvin (1970). . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.2 Turbulent length scale LE variation of a round jet. (a) Schematic
illustration. (b) Self-similar RMS turbulent velocity profile esti-
mated using Eq. 5.7, Chapter 5. uc is the jet centerline mean
velocity. (c) Turbulent length scale normalized with Gaussian jet
half-width bg = 0.114x, estimated using Eqs. 5.7 and 5.8, Chapter
5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.3 Motion of sediment particle in fluid. xp is the particle position; up
is particle velocity; uf is fluid velocity at the location of particle;
ur is relative velocity. . . . . . . . . . . . . . . . . . . . . . . . . . 36

xi
3.4 Time history of turbulent velocity (w ) and autocorrelation (R)
of a single realization, t = 0.001s. LE = 0.005m, = 0.01m/s,
TE = 0.5s, ws = 0.01m/s. . . . . . . . . . . . . . . . . . . . . . . . 39
3.5 Time history of particle position (z) of a single realization, t =
0.001s. LE = 0.005m, = 0.01m/s, TE = 0.5s, ws = 0.01m/s. . . . 40
3.6 Ten realizations (black lines) of the time history of particle posi-
tion (z), t = 0.001s. LE = 0.005m, = 0.01m/s, TE = 0.5s,
ws = 0.01m/s. Red thick line is constant settling velocity. . . . . . 41
3.7 Comparison between analytical (Eqs. 3.39 and 3.41) and numerical
solutions of particle falling in stagnant fluid, d = 50m, p = 2500
kg/m3 , f = 1000 kg/m3 , ws = 2.044mm/s . . . . . . . . . . . . . 44
3.8 Comparison of numerical model prediction and experimental data
of particle falling in stagnant fluid (Brush et al. , 1964). s =
p /f = 2.5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
3.9 Comparison of the numerical model and the experimental results
of Ho (1964) of spheres settling in homogeneously vertically oscil-
lating fluid. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.10 The particle and fluid velocities, and forces acting on the particle.
Rep = 230, s = 2.65, T = 0.2s, wf 0 = 2m/s. . . . . . . . . . . . . . 48
3.11 Schematic representation of a hypothetical experiment of dropping
sediment particles in homogeneous turbulence. . . . . . . . . . . . 49
3.12 Cumulative probability distribution of model predicted sediment
settling velocities under different /ws with AE = 1 and LE =
0.005m, (a) glass particles, (b) plastic particles. . . . . . . . . . . 52
3.13 The predicted apparent settling velocity compared with the ex-
perimental data of settling particle in grid turbulence of Murray
(1970). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
3.14 The predicted apparent settling velocity compared with the ex-
perimental data of settling particle in grid turbulence of Zhou and
Cheng (2009) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
3.15 The predicted apparent settling velocity for the particles used in
this experiment, (a) without Basset force, (b) with Basset force. . 54
3.16 The predicted apparent settling velocity, best-fitted with Eq. 3.44,
and compared with the experimental data of settling particle in
grid turbulence of Murray (1970), Nielsen (1993) and Zhou and
Cheng (2009) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
3.17 A downward sediment laden jet. . . . . . . . . . . . . . . . . . . . 57
3.18 Predicted sediment concentration (C/C0 ) of vertical sediment laden
jet, cases of Singamsetti (1966), (a) 68m, (b) 230m, (c) 460m
and (d) analytical pure jet tracer concentration. . . . . . . . . . . 58

xii
3.19 Comparison of centerline sediment concentration for experimental
cases of Singamsetti (1966). Model prediction with Basset force
is carried out on cases of 68, 230 and 460 m, with very little
difference with the one without Basset force. . . . . . . . . . . . . 59
3.20 Comparison of cross-sectional sediment concentration for experi-
mental cases of Singamsetti (1966). Model prediction with Basset
force is carried out on cases of 68, 230 and 460 m, with very little
difference with the one without Basset force. . . . . . . . . . . . . 60
3.21 Comparison of predicted and measured centerline sediment mass
flux for experimental cases of Parthasarathy and Faeth (1987) . . 62
3.22 Comparison of predicted and measured cross-sectional sediment
sediment mass flux for experimental cases of Parthasarathy and
Faeth (1987) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

4.1 Experimental set up of horizontal sediment-laden jet. . . . . . . . 66


4.2 The two types of hourglass for feeding sediment in the experiment:
(a) for glass particles, (b) for plastic particles (IP3 and MF). . . . 69
4.3 Detailed internal dimension of hourglasses . . . . . . . . . . . . . 70
4.4 Time-mass discharge relationships of hourglasses for (a) IP3 par-
ticles and (b) MF particles. . . . . . . . . . . . . . . . . . . . . . 71
4.5 The two types of sediment collection trays for measuring: (a) 1D,
(b) 2D deposition profiles. . . . . . . . . . . . . . . . . . . . . . . 73
4.6 Glass vials for collecting sediment for weighing (diameter = 4cm;
height = 5cm). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
4.7 Calibration of sediment concentration measurement by mass bal-
ance between the jet cross section and bottom deposition. . . . . . 82
4.8 Comparison of measured and analytical axial centerline velocity
uc (Eq. 4.9). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
4.9 Comparison of measured and analytical cross-sectional axial cen-
terline velocity u(x, r)/uc (x), Eq. 4.10. . . . . . . . . . . . . . . . 84

4.10 Axial RMS turbulent fluctuations u u /uc . . . . . . . . . . . . . 85

4.11 Radial (vertical) RMS turbulent fluctuations w w /uc . . . . . . . 85
4.12 Measured time history of turbulent velocity fluctuations (u , w )
and its local acceleration (u /t, w /t). u0 = 0.39 m/s (Qj =
40 L/h), x = 30D, z = 0, u = 0.080 m/s, w = 0.0014 m/s.
Measurement interval = 0.01s. . . . . . . . . . . . . . . . . . . . . 87
4.13 The correlation between |u | and |u /t|, |w | and |w /t|, at jet
centerline (z = 0). . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
4.14 The correlation between |u | and |u /t|, |w | and |w /t|, along
vertical transects at x = 30D and x = 50D . . . . . . . . . . . . . 89

5.1 Experiment of horizontal sediment-laden jets . . . . . . . . . . . . 94

xiii
5.2 Visualization of case u0 = 0.786 m/s, (a) sediment laden, (b) pure
jet. The dashed line is the jet centerline; the dash-dotted line is
the top-hat width of the jet, defined by bT = 0.16x. . . . . . . . . 96
5.3 Visualization of case u0 = 0.492 m/s, (a) sediment laden, (b) pure
jet. The dashed line is the jet centerline; the dash-dotted line is
the top-hat width of the jet, defined by bT = 0.16x. . . . . . . . . 97
5.4 Visualization of IP3J80, u0 = 0.786 m/s, plastic IP3 particle d50 =
716m , ws = 2.2cm/s . . . . . . . . . . . . . . . . . . . . . . . . 98
5.5 Typical 1D sediment deposition profiles (present study and Lee,
2010). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
5.6 Normalized 1D deposition profiles of glass (Lee, 2010) and natural
sand particles (Li, 2006) . . . . . . . . . . . . . . . . . . . . . . . 101
5.7 Normalized 1D deposition profiles of plastic IP3 particles (present
study) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
5.8 Normalized 1D deposition profiles of melamine MF particles (present
study) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
5.9 Normalized 1D deposition profiles (fitted equations, Eq. 5.2) of
different particles . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
5.10 2D deposition profiles of G180 particles (present study). Contours
in g/m2 /s. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
5.11 2D deposition profiles of IP3 particles (present study). Contours
in g/m2 /s. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
5.12 CFD predicted and observed (Li, 2006) longitudinal particle depo-
sition pattern of four experiments using sand particles. The solid
lines represent the prediction using the adjusted settling velocity:
0.75ws whole field adjustment; wsa local adjustment according to
/ws . The dashed line is prediction using the stillwater settling
velocity ws . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
5.13 CFD predicted and observed (Lee, 2010) longitudinal particle de-
position pattern of six experiments using spherical glass particles.
The solid lines represent the prediction using the adjusted settling
velocity: 0.75ws whole field adjustment; wsa local adjustment ac-
cording to /ws . The dashed line is prediction using the stillwater
settling velocity ws . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
5.14 (a) Ratio of RMS turbulent velocity to settling velocity (/ws ),
and (b) ratio of apparent settling velocity to terminal settling
velocity (wsa /ws ), in different cross-section of a jet, Jet case FJ42:
deq = 133m, u0 = 0.41m/s, ws = 1.41cm/s . . . . . . . . . . . . . 108
5.15 CFD predicted and observed (Lee, 2010) longitudinal particle de-
position pattern of two experiments using plastic particles. 0.65ws , 0.75ws
whole field uniform adjustment of stillwater settling velocity ws ;
wsa local adjustment according to /ws . . . . . . . . . . . . . . . 108

xiv
5.16 Theoretical and CFD predicted jet mean velocity profiles, (a) Lon-
gitudinal velocity, Eq. 5.5, (b) transverse velocity (positive out-
wards), Eq. 5.6. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
5.17 Turbulent velocity fluctuation and dissipation rate dissipation,
predicted by CFD model and fitted with empirical equations, (a)
RMS turbulent velocity fluctuation, Eq. 5.7, (b) Turbulent energy
dissipation rate, Eq. 5.8. . . . . . . . . . . . . . . . . . . . . . . . 111
5.18 Comparison of measured and particle tracking predicted deposi-
tion rate profiles of experiments using natural sand particles (Li,
2006). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
5.19 Comparison of measured and particle tracking predicted depo-
sition rate profiles of experiments using spherical glass particles
(Lee, 2010). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
5.20 Comparison of measured and particle tracking predicted deposi-
tion rate profiles of experiments using spherical plastic particles
(IP3). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
5.21 Comparison of measured and particle tracking predicted deposi-
tion rate profiles of experiments using plastic particles (Lee, 2010). 116
5.22 Comparison of measured and particle tracking predicted deposi-
tion rate profiles of experiments using melamine particles. . . . . . 117
5.23 Observed and predicted (particle tracking) deposition pattern of
the experiment CJ58 (Li, 2006), deq = 166m, u0 = 0.57m/s,
ws = 1.97m/s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
5.24 Observed and predicted deposition profile of the experiment with
G215 particles, d50 = 215m, ws = 2.65m/s. Contour in g/m2 /s.
The jet nozzle is located at x = 0, y = 0. . . . . . . . . . . . . . . 120
5.25 Observed and predicted deposition profile of the experiment with
G180 particles, d50 = 180m , ws = 1.98m/s. Contour in g/m2 /s.
The jet nozzle is located at x = 0, y = 0. . . . . . . . . . . . . . . 121
5.26 Observed and predicted deposition profile of the experiment with
G115 particles, d50 = 115m , ws = 1.00m/s. Contour in g/m2 /s.
The jet nozzle is located at x = 0, y = 0. . . . . . . . . . . . . . . 122
5.27 Observed and predicted deposition profile of the experiment with
IP3 particles, d50 = 716m , ws = 2.2m/s. Contour in g/m2 /s.
The jet nozzle is located at x = 0, y = 0. . . . . . . . . . . . . . . 123
5.28 Observed and predicted deposition profile of the experiment with
MF particles, deq = 347m , ws = 2.2m/s. Contour in g/m2 /s.
The jet nozzle is located at x = 0, y = 0. . . . . . . . . . . . . . . 124
5.29 Sensitivity tests on the Basset, added mass and fluid acceleration
forces. 1D deposition profiles. G: gravity; D: drag force; F: fluid
acceleration; A: added mass; B: Basset. (a)-(c): Glass (Lee, 2010);
(d): Sand (Li, 2006); (e)-(f): plastic, present study . . . . . . . . 126

xv
5.30 Sensitivity tests on the Basset force (present experiments). 2D
deposition profiles. Contour in g/m2 /s. The jet nozzle is located
at x = 0, y = 0. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
5.31 Measured and particle tracking predicted (simplified model) cross-
sectional sediment concentration (normalized with the maximum),
Case G180J80 (Lee, 2010), u0 = 0.78m/s, ws = 2.06cm/s, lm /D =
32.9. Dashed circle represents the top-hat profile of the jet. . . . . 130
5.32 Measured and particle tracking predicted (simplified model) cross-
sectional sediment concentration (normalized with the maximum),
Case G180J50 (Lee, 2010), u0 = 0.49m/s, ws = 2.05cm/s, lm /D =
20.8. Dashed circle represents the top-hat profile of the jet. . . . . 131
5.33 Measured and particle tracking predicted (simplified model) cross-
sectional sediment concentration (normalized with the maximum),
Case G215J60 (Lee, 2010), u0 = 0.59m/s, ws = 2.69cm/s, lm /D =
18.9. Dashed circle represents the top-hat profile of the jet. . . . . 132
5.34 Measured and particle tracking predicted (simplified model) cross-
sectional sediment concentration (normalized with the maximum),
Case G115J70 (Lee, 2010), u0 = 0.68m/s, ws = 1.03cm/s, lm /D =
57.6. Dashed circle represents the top-hat profile of the jet. . . . . 133
5.35 Measured and particle tracking predicted (full model) cross-sectional
sediment concentration (normalized with the maximum), Case
IP3J80 (present experiment), u0 = 0.78m/s, ws = 2.02cm/s,
lm /D = 34.5. Dashed circle represents the top-hat profile of the jet.134
5.36 Measured and particle tracking predicted (full model) cross-sectional
sediment concentration (normalized with the maximum), Case
IP3J50 (present experiment), u0 = 0.49m/s, ws = 2.02cm/s,
lm /D = 21.5. Dashed circle represents the top-hat profile of the jet.135
5.37 Measured and particle tracking predicted centerline maximum con-
centration, G215 particles, ws = 2.69cm/s (Lee, 2010). Dashed
line is the theoretical tracer concentration variation for a free jet. 136
5.38 Measured and particle tracking predicted centerline maximum con-
centration, G180 particles, ws = 2.06cm/s (Lee, 2010). Dashed
line is the theoretical tracer concentration variation for a free jet. 136
5.39 Measured and particle tracking predicted centerline maximum con-
centration, G115 particles, ws = 1.03cm/s (Lee, 2010). Dashed
line is the theoretical tracer concentration variation for a free jet. 137
5.40 Measured and particle tracking predicted centerline maximum con-
centration, IP3 particles, ws = 2.02cm/s (present experiments).
Dashed line is the theoretical tracer concentration variation for a
free jet. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137

6.1 A general schematic representation of a sediment laden buoyant jet.142


6.2 The JETLAG model (Lee and Cheung, 1990; Lee and Chu, 2003). 145

xvi
6.3 The Gaussian variation of mean axial velocity and the equivalent
top-hat profile. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
 1
6.4 Comparison of theoretical ucu(x)
0
= 6.2 Dx and JETLAG model
predicted centerline velocity for a pure jet (u0 = 0.786m/s, D =
6mm), showing the effect of inclusion of the zone of flow estab-
lishment (ZFE). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147

6.5 CFD model predicted turbulent intensity and dissipation rate in


the jet and plume regime. F rl is the local densimetric Froude
number defined as Eq. 6.6. Solid symbols: plume regime; open
symbols: jet regime. . . . . . . . . . . . . . . . . . . . . . . . . . 147

6.6 (a) The external flow field of a horizontal buoyant jet (u0 =
0.65m/s, F r = 19.5) and (b) its influence on particle velocity.
Note the different vector scales . . . . . . . . . . . . . . . . . . . . 150

6.7 The point sink approach for determining the external flow (Lai,
2009). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151

6.8 Comparison of the radial velocity using the point sink method (as-
suming a solid boundary for the plane x = 0) with the analytical
solution (Eq. 6.13). . . . . . . . . . . . . . . . . . . . . . . . . . . 151

6.9 The spreading current after jet impingement at the free surface. . 153

6.10 Comparison of integral model and CFD predicted mean spreading


current velocity us , thickness hs and mean reduced gravity gs . . . 155

6.11 The two fall out mechanisms from vertical sediment-laden buoyant
jet. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158

6.12 Comparison of model prediction of bottom deposition of vertical


sediment-laden jets and experimental data of Neves and Fernando
(1995). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161

6.13 Relation between lm and the maximum height of rise zm , model


prediction for vertical sediment-laden jet experiments of Neves and
Fernando (1995). . . . . . . . . . . . . . . . . . . . . . . . . . . . 162

6.14 Relation of the maximum deposition rate Fmax and position rmax

with lm in vertical sediment-laden jet experiments. The constant


of 3.0 and 0.1 are compared with the value of 1.4 and 0.15 given
by Neves and Fernando (1995) (Eqs. 6.31 and 6.32. . . . . . . . . 162

6.15 Comparison of predicted and measured radial sediment deposi-


tion profile of the four vertical sediment-laden jet cases of Ernst
et al. (1996). r is the radial distance from the nozzle. F/Fmax
is the deposition (g/m2 /s) normalized against the maximum pre-
dicted/measured values. . . . . . . . . . . . . . . . . . . . . . . . 164

xvii
6.16 Comparison of predicted and measured radial sediment deposition
profile of the four vertical sediment-laden plume cases of Ernst et
al. (1996). r is the radial distance from the nozzle. F/Fmax is
the deposition (g/m2 /s) normalized against the maximum pre-
dicted/measured values. . . . . . . . . . . . . . . . . . . . . . . . 164
6.17 Comparison of numerical model prediction and experimental data
of vertical sediment jet experiments: (a)-(c) Sparks et al. (1991),
(d) Zarrebini and Cardoso (2000). r is the radial distance from
the nozzle. F/Fmax is the deposition (g/m2 /s) normalized against
the maximum predicted/measured values. . . . . . . . . . . . . . 166
6.18 Comparison of the Full Model (with and without Basset force) and
Simplified Model for vertical sediment jet experiments of (a)-(b):
Neves and Fernando (1995) (p = 1040 kg/m3 ); (c)-(f): Ernst et
al. (1996) (p = 3210 kg/m3 ). . . . . . . . . . . . . . . . . . . . . 168
6.19 Comparison of CFD predicted and measured longitudinal deposi-
tion profile (g/m/s) of Li (2006)s horizontal sediment-laden buoy-
ant jet experiments. . . . . . . . . . . . . . . . . . . . . . . . . . . 173
6.20 CFD predicted sediment concentration and conservative tracer
concentration for buoyant jets, coarse sand experiments. The
dashed line is the jet trajectory and top-hat boundary. . . . . . . 174
6.21 CFD predicted sediment concentration and conservative tracer
concentration for buoyant jets, fine sand experiments. The dashed
line is the jet trajectory and top-hat boundary. . . . . . . . . . . . 175
6.22 Comparison of particle tracking model predicted and measured
longitudinal deposition profile (g/m/s) of Li (2006)s horizontal
sediment-laden buoyant jet experiments, Coarse Sand. . . . . . . . 178
6.23 Comparison of particle tracking model predicted and measured
longitudinal deposition profile (g/m/s) of Li (2006)s horizontal
sediment-laden buoyant jet experiments, Fine Sand. . . . . . . . . 178
6.24 Predicted 2D deposition profiles for (a) CB66 and (b) FB58. . . . 179
6.25 Examples of particle trajectory in case FB66 with fine sediment.
The dash-dotted line is the jet centerline; dashed lines are the jet
top-hat boundaries . . . . . . . . . . . . . . . . . . . . . . . . . . 180
6.26 Comparison of particle tracking model predicted and measured
longitudinal deposition profile (g/m/s) of Lee (2010)s horizontal
sediment-laden buoyant jet experiments, S199 particles. . . . . . . 180
6.27 Comparison of particle tracking model predicted and measured
longitudinal deposition profile (g/m/s) of Lee (2010)s horizontal
sediment-laden buoyant jet experiments, S153 particles. . . . . . . 181
6.28 Comparison of particle tracking model predicted and measured
longitudinal deposition profile (g/m/s) of Lee (2010)s horizontal
sediment-laden buoyant jet experiments, G215 particles. . . . . . 182

xviii
6.29 Comparison of particle tracking model predicted and measured
longitudinal deposition profile (g/m/s) of Lee (2010)s horizontal
sediment-laden buoyant jet experiments, G180 particles. . . . . . 183
6.30 Comparison of particle tracking model predicted and measured
longitudinal deposition profile (g/m/s) of Lee (2010)s horizontal
sediment-laden buoyant jet experiments, G115 particles. . . . . . 184
6.31 JETLAG predicted jet trajectories and boundary profiles. . . . . . 184
6.32 Comparison of particle tracking model predicted and measured
longitudinal deposition profile (g/m/s) of Lee (2010)s horizontal
sediment-laden buoyant jet experiments, plastic particles (p =
1.16 g/cm3 ). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186
6.33 Comparison of model predicted and measured 1D and 2D deposi-
tion profiles Cuthbertson and Davies (2008)s horizontal sediment-
laden buoyant jet experiments, plastic particles (p = 1.5 g/cm3 ).
The 1D deposition profiles are normalized by the sediment mass
flux at the jet nozzle, in the unit of m1 . . . . . . . . . . . . . . . 187

7.1 Visualization of a vertical sediment-laden jet discharging upwards


- Example 1. The box is for visualization only. Free lateral bound-
ary is used in the model. . . . . . . . . . . . . . . . . . . . . . . . 193
7.2 Predicted deposition profile of vertical sediment-laden jet discharg-
ing upwards - Example 1. The jet and spreading current profiles
are also indicated. . . . . . . . . . . . . . . . . . . . . . . . . . . . 194
7.3 (a) The predicted jet trajectory of Wah Fu Outfall - prototype.
(b) Experimental image of a 1:11 scale model (Lee and Chu, 2003) 195
7.4 Predicted deposition profiles of horizontal sediment-laden buoy-
ant jet - Wah Fu Outfall: (a) 1D (in tonnes/m/yr); (b) 2D (in
tonnes/m2 /yr) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196
7.5 Predicted jet trajectory of Wah Fu Outfall in a coflowing ambient
current of 0.03m/s and a visualization of the predicted sediment
distribution on the centerline plane. . . . . . . . . . . . . . . . . . 198
7.6 Predicted deposition profiles of horizontal sediment-laden buoyant
jet - Wah Fu Outfall with an ambient current of 0.03m/s: (a) 1D
(in tonnes/m/yr); (b) 2D (in tonnes/m2 /yr) . . . . . . . . . . . . 199

A.1 Numerical integration of the Basset force term. Velocities at and


before time t is known and the velocity at time t + t is to be
predicted. The Basset integral is to be evaluated from 0 to t + 21 t.212
A.2 Numerical solution of a particle falling in stagnant ambient, with
the Basset force term. d = 50m, p = 2500 kg/m3 , f = 1000
kg/m3 , ws = 2.0437mm/s, t = 104 s. (a) Time history of
particle velocity and acceleration; (b) the Basset integrand at
t = 0.002, 0.005, 0.01s. The particle velocity wp is normalized by
the stillwater settling velocity ws . . . . . . . . . . . . . . . . . . . 213

xix
B.1 Computation mesh for horizontal sediment-laden jets . . . . . . . 218
B.2 CFD predicted velocity and tracer concentration (C/C0 ) field of
x z plane at y = 0. u0 = 0.41 m/s; D = 6 mm; C0 = 5.29 g/L.
The dashed lines are the jet top-hat boundary defined as bT = 0.16x.220
B.3 CFD predicted tracer concentration field (C/Cc ) of y z plane at
x = 10D and x = 30D. u0 = 0.41 m/s; D = 6 mm. The dashed
circles are the jet top-hat boundary defined as bT = 0.16x. . . . . 220
B.4 Comparison of CFD predicted and theoretical centerline velocity
uc /u0 = 6.2(x/D)1 . . . . . . . . . . . . . . . . . . . . . . . . . . 221
B.5 Predicted jet Gaussian half width using the realizable k model.
= 0.114 is the commonly adopted value for round jets. . . . . . 221
B.6 CFD predicted horizontal buoyant jet velocities. (i) Initial region,
(ii) bend up region. The dashed lines are the JETLAG predicted
jet boundary. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
B.7 Comparison of computed horizontal buoyant jet trajectories of
CFD and JETLAG. . . . . . . . . . . . . . . . . . . . . . . . . . . 223
B.8 Comparison of CFD computed horizontal buoyant jet centerline
dilutions with Cederwall equation. . . . . . . . . . . . . . . . . . . 223
B.9 Computation mesh for vertical axisymmetric buoyant jet imping-
ing the surface. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224
B.10 CFD predicted spreading layer velocity field and concentration
field (C/C0 ). u0 = 0.351m/s, D = 7mm, F r = 10.0. . . . . . . . . 225
B.11 CFD predicted spreading layer velocity field and concentration
field (C/C0 ). u0 = 0.146m/s, D = 7mm, F r = 4.0. . . . . . . . . . 226

C.1 Density of ethanol-water mixture. . . . . . . . . . . . . . . . . . . 228


C.2 Dynamic viscosity of ethanol-water mixture. . . . . . . . . . . . . 228
C.3 Variation of settling velocity in a horizontal buoyant jet using
ethanol-water mixture as the source of buoyancy. Case FB58 (Li,
2006): u0 = 0.57m/s; F r = 14.5; d = 133m (sand). (a) JETLAG
predicted centerline jet fluid density (c ) and estimated dynamic
viscosity (c ) using Figs. C.1 & C.2. (b) Estimated sediment set-
tling velocity at jet centerline using Soulsby equation. . . . . . . . 229
C.4 Variation of settling velocity in a horizontal buoyant jet using
ethanol-water mixture as the source of buoyancy. Case G180B90Fr17
(Lee, 2010): u0 = 0.88m/s; F r = 18.3; d = 180m (glass). (a)
JETLAG predicted centerline jet fluid density (c ) and estimated
dynamic viscosity (c ) using Figs. C.1 & C.2. (b) Estimated sed-
iment settling velocity at jet centerline using spherical drag law. . 230
C.5 Model comparison of using local settling velocity and ambient set-
tling velocity in the prediction of sediment deposition on buoyant
jets. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231

xx
List of Tables

1.1 Typical sewage effluent quality and Hong Kong Water Quality
Objective (EPD, 2006). . . . . . . . . . . . . . . . . . . . . . . . . 8

3.1 Experiments of Ho (1964) of spheres settling in homogeneously


vertically oscillating fluid. . . . . . . . . . . . . . . . . . . . . . . 46
3.2 Particles used in grid turbulence experiments of Murray (1970).
= 106 m2 /s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
3.3 Particles used in grid turbulence experiments of Zhou and Cheng
(2009) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
3.4 Experiments of Singamsetti (1966) . . . . . . . . . . . . . . . . . 61
3.5 Experiments of Parthasarathy and Faeth (1987) . . . . . . . . . . 61

4.1 Mass feeding rates for the hourglasses . . . . . . . . . . . . . . . . 68


4.2 Properties of particles in the present experiments. . . . . . . . . . 76
4.3 Horizontal momentum jet experiments using glass particles, mea-
suring 2D bottom depositions. . . . . . . . . . . . . . . . . . . . . 79
4.4 Horizontal momentum jet experiments using IP3 particles, mea-
suring 1D and 2D bottom depositions. . . . . . . . . . . . . . . . 79
4.5 Horizontal momentum jet experiments using MF particles, mea-
suring 1D and 2D bottom depositions. . . . . . . . . . . . . . . . 80
4.6 Cross-sectional sediment concentration measurement, IP3 particles. 81
4.7 Time average of the velocity derivatives for the jet case of Q = 50
L/h (u0 = 0.492 m/s). D = jet diameter = 6mm. . . . . . . . . . 91
4.8 Velocity derivatives and the result of correlation for the jet case
of Q = 50 L/h (u0 = 0.492 m/s). D = jet diameter = 6mm. . . . 92

5.1 Summary of Li (2006) and Lee (2010)s horizontal sediment-laden


jet experiments for bottom deposition and cross-sectional sedi-
ment concentration measurement. Jet diameter D = 6mm. lm =
1/2
momentum-settling length scale = M0 /ws . . . . . . . . . . . . . 99
5.2 The inferred particle equivalent diameter deq for non-spherical
particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112

xxi
6.1 Numerical experiments for vertical sediment-laden pure jets (Neves
and Fernando, 1995). . . . . . . . . . . . . . . . . . . . . . . . . . 160
6.2 Vertical buoyant jet experiments (fall out from jet margin) in
Ernst et al. (1996) . . . . . . . . . . . . . . . . . . . . . . . . . . 163
6.3 Vertical buoyant jet experiments (fall out from surface current) of
Sparks et al. (1991), and Zarrebini and Cardoso (2000) . . . . . . 166
6.4 Horizontal buoyant jet experiments of Li (2006). . . . . . . . . . . 170
6.5 Horizontal buoyant jet experiments of Lee (2010), S199 and S153
particles. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
6.6 Horizontal buoyant jet experiments of Lee (2010), G215, G180 and
G115 particles. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
6.7 Horizontal buoyant jet experiments of Lee (2010), plastic particles
(d = 621m , p = 1.16 g/cm3 ). . . . . . . . . . . . . . . . . . . . 185
6.8 Horizontal buoyant jet experiments of Cuthbertson and Davies
(2008), plastic particles (d = 550m , p = 1.5 g/cm3 ). . . . . . . 186

xxii
List of Symbols

TE
AE = , parameter in autocorrelation function
LE
Ap = projected area of particle = d2 /4
b, bg = Gaussian jet half-width
bT = top-hat jet half-width
B = Basset integral
B0 = Q0 g0 , initial jet specific buoyancy flux
Bs = specific radial buoyancy flux of spreading layer
C0 = jet initial sediment concentration
CD = drag coefficient = f (Rep )
CM = added mass coefficient
Cmax = maximum jet cross-sectional sediment concentration
C = 0.09, constant in k turbulence model
d = particle diameter
d50 = median particle diameter for spherical particles
deq = equivalent particle diameter for non-spherical particles
D = Jet diameter
F = Bottom deposition rate in vertical sediment-laden jets,
in g/m2 /s
Fs = transversely lumped (1D) sediment bottom deposition rate,
in g/m/s
Fr = initial jet densimetric Froude number
F rl = local jet densimetric Froude number
g = gravitational acceleration = 9.81 m/s2
g0 = jet initial reduced gravity
gs = reduced gravity for spreading layer
hs = thickness of spreading layer
H = height of release in homogeneous grid turbulence
k = turbulent kinetic energy
lb = particle inertia-buoyancy length scale, Chapter 6

xxiii
lm = momentum-settling length scale
ls = momentum-buoyancy length scale
LE , Lz = turbulent length scale
mi = strength of point sink i
M0 = Q0 u0 , initial jet kinematic momentum flux
Ms = kinematic radial momentum flux of spreading layer
N, Np = number of particles in numerical simulations
Q0 = u0 D2 /4, jet initial flow rate
Qs = volumetric flux of spreading layer
r = radial coordinate
R = velocity autocorrelation function
Ri = Richardson number for spreading layer
Re = jet Reynolds number
Rep = particle Reynolds number
s = specific gravity of particles
s = streamwise coordinate along jet centerline (Chapter 6)
Sct = turbulent Schmidt number in CFD model
t = time
t = time step
T = average settling time of particles in turbulence
TE = Eulerian time scale of turbulence
TL = Lagrangian time scale of turbulence
u0 = initial jet velocity
ua = ambient velocity
uc = jet centerline velocity
us = radial velocity of spreading layer
uT = jet top-hat velocity
uf = (uf , vf , wf ) = fluid velocity
uf = time-mean fluid velocity
u , uf

= turbulent fluid velocity fluctuations
up = (up , vp , wp ) = particle velocity
ur = relative velocity between particle and fluid
vr = mean radial velocity of a jet
Vp = volume of particle = d3 /6
we = interfacial entrainment velocity for spreading layer
ws = sediment settling velocity in stillwater
wsa = apparent sediment settling velocity in turbulence
x, y, z = Cartesian coordinates
xi , yi , zi = location of the ith jet element

xxiv
xp = (xp , yp , zp ), particle position

Greek symbols

, g = jet entrainment coefficient at r = bg


T = jet entrainment coefficient at r = bT
= jet spreading rate
= turbulent energy dissipation rate
= ratio between the jet Gaussian half-width
of concentration and velocity
i = interfacial friction coefficient for spreading layer
= dynamic molecular viscosity of fluid
= kinematic molecular viscosity of fluid = /f
t = turbulent viscosity of fluid
a = ambient fluid density
0 = initial jet fluid density
f = fluid density
p = particle density
= root-mean-square turbulent fluctuation
= dummy time variable in Basset integral
= velocity potential
= randomly generated numbers following a Gaussian
distribution with zero mean and unit variance
= oscillation frequency

Empirical constants

A, B, C Empirical constants in the relation of


wsa and , Chapter 3
A, B, C Empirical constants in the relation of
normalized sediment bottom deposition rate
and normalized longitudinal distance, Chapter 5
C1 C3 Empirical constants in normalized
jet turbulent intensity profile
C4 C6 Empirical constants in normalized
jet turbulent dissipation rate profile

xxv
xxvi
Chapter 1

Introduction

1.1 Background
The turbulent transport of sediment or particulate matters in a fluid flow (water,
air) is a common and important phenomenon. Turbulent jets and plumes (buoy-
ant jets) constitute a major class of turbulent flows. Sediment or particle-laden
buoyant jets are common in natural environment and engineering applications.
Volcanic eruption and black smokers released from deep ocean hydrothermal
vents are examples of particle-laden jets in the natural environment. Sediment-
laden river plumes bring huge amount of sediment to the ocean, shaping the
land-form of coastal/estuarine area. In marine engineering, reclamation and
dumping of dredged spoil and sludge to the ocean is usually carried out in the
form of sediment-laden jets. Submarine discharge of wastewater effluent in the
form of buoyant jets also contains sediment particles.

1.1.1 Volcanic eruption


Volcanoes are prominent geological features on the Earth. They exists on land
and beneath the ocean, releasing hot magma, volcanic ash and gases. Volcanoes
have been shaping the landscape of the Earth since the formation of the Earth
4.6 billion years ago. There are about 1500 active volcanoes in the world and
around 50 of them erupt each year. An estimated 500 million people live near
active volcanoes (European Space Agency, 2009).
Erupting volcanoes can pose many hazards to human life. During a vol-
canic eruption, a mixture of volcanic gases and pyroclastic materials (magma,
suspended crystals, volcanic ashes) are discharged vertically upwards into the
atmosphere in the form of a particle-laden plume. The hot solid materials over
1000 C finally fall back to the Earth, forming a gravity flow of pyroclastic materi-
als that travels down the volcano (Fig. 1.1). Volcanic ashes and pyroclastic flows
can be deadly and disastrous. The ancient city of Pompeii was buried by falling
volcanic ashes from the eruption of Mount Vesuvius in 79 A.D., killing about
20,000 people. In 1902, the pyroclastic flow from the eruption of Mount Pelee
on Martinique, Caribbean Sea, destroyed the town of Saint-Pierre. Only two of

1
its 30,000 residents survived in the disaster. Deposited materials on the ground
mix with water and snow to form debris flow called lahar which is tremendously
destructive. A small eruption of Nevado del Ruiz, Colombia in 1987, produced
large lahars that covered the city of Armero, claiming the lives of 25,000 people.
Suspended volcanic ashes threaten aviation safety as they cause aircraft en-
gines to malfunction. In 1989, a flight carrying 245 passengers lost power to all
four engines after flying into an ash cloud from the erupting Mount Redoubt,
Alaska, luckily landed safely for emergency. In mid April 2010, the spreading
of volcano ashes from a volcanic eruption in Iceland (about 1800km from Lon-
don) resulted in the suspension of flights in northern Europe for a week and
resulted tremendous economic losses. It is important to understand and predict
the transport and deposition of volcanic materials from an eruption in order to
safeguard human lives and reduce economic loss.

1.1.2 Submarine hydrothermal vents - black smokers

A submarine hydrothermal vent is a fissure in deep ocean bottom from which


geothermally heated water issues. It was first discovered in 1979 in the Eastern
Pacific Ocean, and later found in other places of the worlds ocean. Hydrothermal
vents are commonly located near tectonically and volcanically active places such
as submarine ridges. Seawater seeped into the seabed rocks is heated by the hot
rocks at depth. Heated water emerges from these vents at temperatures ranging
from 60 C up to as high as over 400 C, under the great pressure from the above
sea water (over 2000 m deep). Many of these hydrothermal plumes are black
in colour because of the dissolved sulphuric minerals, hence the name black
smokers (Fig. 1.2).
Submarine hydrothermal vents are locations of high scientific values on biol-
ogy, geology and oceanography. Hydrothermal vents provide energy and nutri-
ents to the unique deep sea organisms. Besides the scientific values, their poten-
tial economic values cannot be overlooked. As the hot water discharges into sea-
water, the mineral-rich discharge cools due to mixing with the low-temperature
ambient ( 2 C) and the mineral precipitates to form a sediment mound around
the vent. These mineral-rich sediment consists of iron, copper, zinc sulphides and
other rare earth minerals with high economic values. Rare earth elements (scan-
dium, yttrium, etc.) are important raw materials for high-technology products
including flatscreen televisions, smart phones and lasers, but they exist sparsely
on the land and are costly to explore and excavate. Recent research showed
that deep-sea mud contains high concentrations of rare-earth elements, possibly
sourced from hydrothermal plumes (Kato et al. 2011). The hydrothermal vent
fields will become an alternative and continuous source of natural mineral supply
other than the limited but diminishing land resources. The scientific research on
the sediment formation and deposition from hydrothermal vents contributes to
the effective exploration of these natural resources with minimal impact to the
ocean environment and ecology.

2
Volcanic plume

Pyroclastic flow

(Source: U.S. Geological Survey, http://volcanoes.usgs.gov/)

Figure 1.1: Eruption of Mayon Volcano, the Philippines in 1984, showing the
volcanic plume of hot gases and volcanic ashes, and streams of pyroclastic flows
running down the slope.

(Source: NOAA, http://www.photolib.noaa.gov/htmls/nur04506.htm)

Figure 1.2: A deep ocean hydrothermal vent - black smoker in Atlantic Ocean.
Precipitated minerals are deposited around the vent.

3
1.1.3 Marine disposal of dredged spoils and sludge

Dredging is necessary for the maintenance of navigation channels, flood allevi-


ation and the supply of sand for land reclamation and other construction ac-
tivities. Dredging spoils have to be properly disposed of as they may contain
various pollutants. Sewage treatment produces organic and bacteria-rich sludge.
Disposal of dredged spoils and sewage sludge in landfill takes up the precious
land resources. They are often disposed of by dumping in the coastal/marine
environment. The waste materials are transported to designated sites by a barge,
and released to the sea in the form of sediment thermal or sediment-laden jets
(Fig. 1.3). Polluted wastes are capped with a thick layer of clean sediments to
isolate them from the marine environment and protect them from storm erosion.
The dumping of waste to the sea can adversely impact the aquatic life. Ma-
rine dumping of contaminated sediment may cause increased turbidity, reduced
dissolved oxygen level and increased organic and heavy metal pollutants. There
are currently four marine disposal sites in Hong Kong waters (Fig. 1.4). The
dumping sites are usually close to sensitive receivers such as marine parks and
fish culture zones. One of the dumping pit for contaminated dredged spoils at
the East of Sha Chau, is within the habitat of the endangered Chinese White
Dolphin. The dumping operation has to be carefully controlled to minimize the
loss of contaminated materials to the ambient, with regular monitoring on water
quality. Predicting the transport and deposition of disposed spoils and sludge
is necessary to ensure that the operations will not cause unacceptable impact to
the marine environment.

(Source: http://www.abc.net.au/news/)

Figure 1.3: Dumping of dredged spoils from a barge.

4
1.1.4 Wastewater disposal in the marine environment

In coastal cities like Hong Kong, marine waters are important resources for
recreation, fish culturing, water supply (for toilet flushing, cooling), navigation,
etc. At the same time, domestic or municipal wastewater is often discharged
into coastal waters after treatment, in proximity to sensitive receivers such as
beaches, fish culture zones and marine reserves (Fig. 1.4). Water quality ob-
jectives/standards (WQO) of these sensitive receivers have to be maintained
to protect public health and the marine environment (see Table 1.1 for Hong
Kong WQO). Water quality parameters of raw sewage often exceeds the WQOs
(Table 1.1), thus proper design of sewage treatment/discharge is essential to re-
move/dilute for complying with the WQO. Submerged sewage outfalls, typically
in the form of a long diffuser pipe with a number of risers, is commonly used
to discharge wastewater in form of buoyant jets into the water body, due to the
large dilution achieved by active turbulent mixing (Fig. 1.5).
Wastewater contains different kinds of solids in a wide range of sizes (<0.1
m to >3 cm). In untreated domestic sewage, the total solid content is 350-1200
mg/L, in which 100-300 mg/L is suspended solid (Metcalf and Eddy, 1991). The
removal of solids is essential in wastewater treatment as large solids may block
pipes and damage equipment. Subjected to various treatment levels, solids of
different sizes are removed. For example, preliminary screening can remove solids
smallest to 200 m in diameter; around 50% of the total solids can be removed.
In chemically enhanced primary treatment (CEPT), up to 80% of suspended
solids can be removed by adding chemicals such as ferric chloride to enhance the
flocculation and settling of suspended solids. The Stonecutters Island Sewage
Treatment Work (SCISTW) under the Harbour Area Treatment Scheme (HATS)
in Hong Kong is the largest CEPT plant in the world, occupying 100,000 m2 of
land and currently treating 1.4 million m3 /d sewage collected from the urban
area of Hong Kong Island and Kowloon via deep tunnels (Fig. 1.4). Nevertheless,
the land scarcity problem in many densely-populated cities prevents the use of
extensive land-based sewage treatment. High-level sewage treatment requires
high cost to build and operate for small communities and developing countries.
Preliminary treatment is still adopted due to its relatively small land occupation
and lower cost. In Hong Kong, there are 11 marine outfalls discharging 0.73
million m3 /d preliminarily treated sewage in 2012 (Fig. 1.4). Eight of them will
be abondoned under HATS Stage 2A in 2014 with sewage transferred to SCISTW
for centralized treatment. Some of these outfalls have been in operation since the
1970-80s and discharged a susbstantial amount of sewage sludge to the marine
environment. The cumulative and long-term impacts of these sediment needs
further assessment.
The environmental and ecological impacts of settled particulate matter are
multi-fold: (1) formation of sediment/sludge banks at the vicinity of an outfall.
Anaerobic decomposition of organics in the deeper sediment layer forms gases
such as hydrogen sulphide which may result in odour and foaming at water
surface (Thomann and Mueller, 1987). (2) Dissolved oxygen demand is exerted
due to the oxidation of organics in the sediment/water interface and may cause

5
severe dissolved oxygen depletion in a stratified water body. (3) Enrichment and
accumulation of harmful chemicals such as heavy metals, carcinogenic substances
(e.g. chlorination byproducts, DDT, PCBs) and endocrine disrupting compounds
(EDCs) as they tend to be stably adsorbed on particle surface and create far-
reaching effects to the ecology system through bio-accumulation in the food web.
Worldwide there is an emerging concern on the environmental impact of en-
docrine disrupting compounds (EDCs) and pharmaceuticals (e.g. antibiotics).
Many chemicals used as ingredients or additives in a number of household prod-
ucts such as personal health products and cosmetics, are found as EDCs. Labora-
tory studies have revealed that a number of these chemicals are capable of causing
reproductive disorders in animals, disrupting their hormone system and result-
ing in cancers, under concentrations as low as the order of 0.1 ng/L (Gomes and
Lester, 2003). Medical evidence on their impact on human health is emerging.
The overuse of antibiotics causes bacterial pathogen resistance. These chemi-
cals are inevitability discharged to marine waters with treated sewage. Minh et
al. (2009) found that antibiotics are ubiquitous in sea water throughout Victoria
Harbour in Hong Kong and sewage discharge is one of the major sources. EDCs
and antibiotics have also been found to adsorb on marine sediment, providing
ground for their accumulation (e.g. Gomes and Lester, 2003; Xu and Li, 2010).
Their impact to aquatic organisms and human beings may last for years.

1.2 Motivation and objective of this study

Sediment-laden buoyant jets have many applications in science and engineering.


The understanding on sediment/particle jet physics is essential for the safeguard
of human lives, the protection of environment and economic development. Math-
ematical modelling and prediction for the mixing and deposition of sediment jets
is required to achieve these purposes.
Single-phase buoyant jets has been extensively studied with robust models
developed for engineering design, prediction and assessment purposes. However
there has been limited research on sediment-laden buoyant jets. Turbulence-
particle interaction in buoyant jets is not well understood. There is hitherto no
general model for predicting the mixing and deposition of a sediment-laden jet
for the impact assessment of sediment/sludge discharged with sewage.
The objective of the present study is to develop a general model to pre-
dict the sediment mixing and deposition of dilute sediment-laden buoyant jets
in arbitrary inclination in stagnant and unstratified ambient (Fig. 1.6). The
model framework is based on three-dimensional (3D) Lagrangian stochastic par-
ticle tracking. Laboratory experiments is conducted to understand the physics
of sediment-laden jets and to acquire data for model validation. 3D computa-
tional fluid dynamics (CFD) modelling is used to study the turbulence-particle
interaction in sediment-laden jets.

6
Sewage outfalls - screening/preliminary
Sewage outfalls - primary/CEPT
Sewage outfalls - secondary
Beaches
SHENZHEN
Fish Culture Zones
Marine Dumping Sites
Marine Parks/Reserves
2230'N
Pearl River New Territories
Estuary

HONG KONG

Sha
Chau
2220'N SCISTW
Kowloon

HATS

Lantau Island Hong Kong Island

2210'N
0 3 6 12 km

11350'E 1140'E 11410'E 11420'E

Figure 1.4: Sewage outfalls in Hong Kong (preliminary, primary and secondary
treatment), marine dumping sites and sensitive receivers (bathing beaches, fish
culture zones and marine reserves). The solid line is the transfer tunnels of HATS
Stage 1 in operation. The dashed line is the transfer tunnels of HATS Stage 2A
commencing in 2014.

Figure 1.5: 3D depiction of the diffuser of Harbour Area Treatment Scheme


(HATS) outfall (1.2km long, 24 rosette risers, each with 8 jets). Two of the
rosette risers are shown in detail.

7
Table 1.1: Typical sewage effluent quality and Hong Kong Water Quality Objec-
tive (EPD, 2006).
Treatment level BOD5 Suspended Unionized E.coli
Solid Ammonia
(mg/L) (mg/L) (mg/L) (count/100mL)
Raw sewage 220 220 20 107
After preliminary 220 200 20 107
treatment
After primary 120 80 20 106 107
treatment
After CEPT 50 40 20 106 107
After secondary 20 20 3 5 104
treatment
HK Water - not to raise Annual mean a Annual geomean
Quality natural ambient 0.021 610
Objective level b Geomean of 5

(0.5-50 mg/L) recent meas.


by 30% 180
a Secondary contact and fish culture zones
b Bathing beaches

Figure 1.6: Experimental image of a sediment-laden jet in this study. Jet velocity
= 0.79m/s; glass particles of 180 m and settling velocity 2.0 cm/s are used.

8
1.3 Outline of the present work
Chapter 2 reviews the physics governing pure and sediment-laden buoyant jets.
Previous studies on sediment-laden jets and plumes are reviewed in detail. The
physics governing the motion of particles, in both stagnant and turbulent fluid
is discussed, followed by a summary on the Eulerian and Lagrangian modelling
of two-phase flows.
Chapter 3 presents the development of a three-dimensional (3D) stochastic
particle tracking model. Two particle tracking models are developed: the Full
Model solves the equation of particle motion for particle velocity and the Sim-
plified Model assumes the particle velocity is a sum of the fluid velocity and the
sediment stillwater settling velocity. The core of both models are a Lagrangian
velocity autocorrelation function which models the loitering and trapping of par-
ticle in turbulent eddies. The particle tracking models are validated against
previous basic experiments.
Chapter 4 provides the details on the experimental investigations performed
for horizontal sediment-laden jets. Bottom deposition and cross-sectional sedi-
ment concentration are measured. Jet turbulent velocity measurement is carried
out to support the assumption in the autocorrelation function in the particle
tracking model.
Chapter 5 summarises the experimental observation and findings of horizontal
sediment-laden momentum jets. 3D Computational fluid dynamics (CFD) mod-
elling is utilized to investigate the turbulence-particle interaction in horizontal
momentum jets. The particle tracking model based on the theory in Chapter 3
is developed to predict sediment deposition and concentration and extensively
validated with experimental measurement.
Chapter 6 presents the development of a general particle tracking model for
arbitrarily inclined sediment-laden buoyant jets in stagnant ambient. It is essen-
tially an integration of the mathematical models for mean jet flow, jet-induced
external current and surface spreading current. The model is validated exten-
sively against previous experiments of vertical and horizontal sediment-laden
buoyant jets.
Chapter 7 summarises the contributions of the present study; suggestions for
future research areas and topics are given.

9
10
Chapter 2

Literature Review

2.1 Introduction
In this thesis, the focus will be on sediment-laden buoyant jets in unstratified
and stagnant fluid (without any cross flow). The study of turbulent sediment-
laden buoyant jets requires the understanding of the physics of turbulent buoyant
jets and turbulence-particle interactions. In this chapter, previous studies of
single-phase turbulent buoyant jet are firstly reviewed, followed by the review
on particle-laden jets, including experimental investigations, and theoretical and
numerical modelling. Secondly, the physics governing the motion of particles, in
both stagnant and turbulent fluid is discussed. The modelling of particle motion
in a turbulent flow is critical to the successful modelling of sediment-laden jets.
Previous studies on particle motion in a turbulent flow is reviewed.

2.2 Single-phase buoyant jet in stagnant fluid


A buoyant jet is an important turbulent flow commonly encountered in engineer-
ing applications and natural environment. As fluid exits the nozzle, turbulence
is induced by shearing with the ambient fluid leading to entrainment and mix-
ing with the environment. A non-vertical buoyant jet gradually bends upwards
under the increased vertical momentum due to buoyant acceleration (Fig. 2.1).

2.2.1 The mean flow properties


The dynamics of a turbulent buoyant jet in stagnant fluid is governed by its
initial kinematic momentum flux M0 = Q0 u0 = u20 D2 /4 and specific buoyancy
flux B0 = Q0 g0 = g0 u0 D2 /4. As the cross-sectional dimension of a jet is much
smaller than the stream-wise dimension, it can be considered a boundary-layer
type problem. The jet becomes completely turbulent after a short distance from
the nozzle, named as the zone of flow establishment (ZFE), inside which the
turbulence has not penetrated into the potential core. In the zone of established
flow (ZEF), the cross-sectional distributions of time-mean velocity and buoyancy

11
External jet
entrainment
induced flow

Figure 2.1: A single phase buoyant jet.

are observed to be self-similar along the jet trajectory. Measured distribution


of velocity and tracer concentration can be well described by Gaussian profiles.
Morton et al. (1956) introduced the entrainment hypothesis for turbulent closure
to obtain analytical solutions of mean velocity and concentration for pure jets
or plumes in stagnant ambient. Alternatively, a jet spreading hypothesis can be
used to for turbulent closure. The time-mean velocity and concentration prop-
erties of pure round jet and plume have been extensively studied (e.g. Fischer et
al. 1979; Woods et al. 1993).
The governing Reynolds-averaged Navier-Stokes equations for fluid flow can
be numerically solved with a turbulence closure model which adopts the Boussi-
nesq eddy viscosity assumption. The turbulent shear stresses are related to the
mean flow gradient with a proportionality factor termed as the turbulent viscos-
ity/diffusivity. The turbulent viscosity/diffusivity is a not a constant and has
to determined from mean flow properties. A commonly adopted model is the
k model (Rodi, 1984), where k and refer to the kinetic energy of turbulence
and its dissipation rate respectively. However the turbulence model may not
necessarily give satisfactory prediction of complex turbulent buoyant flows.
In engineering application, integral model is considered to be a robust and
efficient way for the prediction of a free buoyant jet. With the self similarity
assumption, the governing equations are integrated over the jet cross sections to
obtain the integral equations in terms of jet volume, momentum and buoyancy
fluxes. The jet trajectory as well as the mean jet properties can be predicted with
conservation of momentum and buoyancy, using either entrainment or spreading
hypotheses. The JETLAG model (Lee and Cheung, 1990; Lee and Chu, 2003)
is a comprehensive and well-validated model for predicting buoyant jets over
a wide range of ambient conditions. For stagnant condition, the unknown jet
trajectory and mean flow properties of an arbitrary inclined buoyant jet can be

12
well-predicted by an integral model using the shear entrainment hypothesis.

2.2.2 Jet/plume turbulence

Turbulence is induced with shearing of the jet flow with the stagnant ambient,
giving rise to jet entrainment and mixing. A number of experimental studies
have been carried out to understand jet and plume turbulence. Wygnanski and
Fielder (1969) measured air jet turbulence using hot-wire anemometry (Re =
Jet Reynolds number = u0 D/ 105 ; Papanicolaou and List (1988) measured
water jet and plume turbulence using Laser Doppler Anemometry (LDA) (Re =
1800 10600); Wang and Law (2002) carried out measurement using Particle
Imaging Velocimetry (PIV) (Re = 150013000). Their study consistently shows
that the maximum root-mean-square (RMS) turbulent velocity is about 25% of
the centerline mean flow velocity in the axial direction and about 20% in the
radial direction. Parthasarathy and Faeth (1987) and Wang and Law (2002)
also showed that in the plume regime, the maximum RMS turbulent velocity
is similar to that of a jet, but the RMS tracer concentration is much higher
(40% of centerline mean concentration compared to 25% for a pure jet). The
cross-sectional distribution of measured RMS turbulent velocity fluctuation is
shown to be self-similar (collapse into the same profile after normalization with
the characteristic jet mean flow velocity and half width) despite the wide range
of Re in the independent experiments. Thus the turbulent quantities can be
deduced from the mean flow properties, for both laboratory (Re 103 104 )
and prototype conditions (Re 105 106 ).

2.2.3 Jet induced external flow

Turbulent entrainment for a buoyant jet creates an irrotational flow external to


the jet (Fig. 2.1, Regime ii). This external flow field is important to the prediction
of sediment particle motion when the particle falls out from the turbulent jet
region. Experimental studies on vertical upward sediment-laden jets and plumes
show that the external flow results in the re-entrainment of particles back to
the jet. However, the external flow field in the previous sediment jet models are
either crudely modelled or neglected.
For their study on multiple jet interaction (Lai and Lee, 2012a), Lai (2009)
developed a general model for predicting the external flow field induced by a
buoyant jet. As far as the outside irrotational flow is concerned, the jet acts as
a distributed entrainment sink. The entrainment flow strength per unit stream-
wise length can be predicted using an integral jet model. The irrotational flow
induced at any given point due to one jet can be predicted as the sum of the
velocities induced by the entrainment sinks. Bottom and free surface boundaries
can be accounted for using the method of images. This greatly contributes to
the modelling of sediment trajectory outside the jet in this study.

13
2.2.4 Gravitational spreading current induced by jet im-
pingement
When a buoyant jet reaches the surface, it spreads radially outwards as gravity
current (Fig. 2.1, Regime iii). The dynamics of the spreading current depends on
the water depth, initial jet configuration and the surface layer thickness. With
experiment and analytical modelling, Lee and Jirka (1981) observed that the
initial thickness of the radial spreading layer is about 0.08 of water depth H
of a vertical buoyant jet in deep water. Kuang and Lee (2006) showed that a
3/4 1/2
stable internal hydraulic jump occurs when ls /H < 3.5, where ls = M0 /B0
is the jet momentum-buoyancy length scale. Akar and Jirka (1994) developed a
model for predicting the gravitational spreading and dilution of ocean outfalls,
accounting for the interfacial and wind-induced mixing. Their modelling idea
is incorporated in the present study for predicting the dynamics of buoyant jet
induced spreading current.

2.3 Sediment-laden jets


2.3.1 Vertical downward sediment-laden jets
Vertically downward sediment jet is a simple case for the study of turbulence-
sediment interaction, as the fluid flow moves in the same direction with the
sediment natural gravitational settling. Singamsetti (1966) is the first to study
the diffusion of sediment in a downward jet using sand particles of different sizes.
Particle concentration across the jet was measured using isokinetic suction. By
estimating the spreading of particle concentration half-width, it was concluded
that sediment diffusion increases with the particle size because of the tendency
of particles to be thrown out of the eddies. However, only the sediment concen-
tration within one Gaussian half width was measured; the estimated spreading
rate may also be prone to measurement error.
Parthasarathy and Faeth (1987) measured simultaneously fluid and particle
velocity, and the particle number flux for a downward vertical sediment-laden jet
in water, using Laser Doppler Anemometry (LDA). Experimental results showed
that the fluid phase is not significantly affected by the presence of sediment
in dilute concentration. Particle velocity does not decay according to x1 like
the liquid due to their finite settling velocity. Numerical Eulerian-Lagrangian
models were compared with experimental data, but the increasing complexity in
modelling the turbulence-particle interaction did not result in predictions that
compare well with data. A simple model regarding the two phase flow as single
phase can also predict sediment-laden jet in water well.
Jiang and Law (2005) reported simultaneous PIV measurement fluid and
particle phase for vertical downward particle-laden jets. Analysis based on two-
phase conservation equations showed that the mean sediment velocity can be
taken as the sum of fluid velocity and the sediment settling velocity. The general
applicability of this assumption has not been studied on horizontal sediment-

14
laden buoyant jets.

2.3.2 Vertical upward sediment-laden jets

There is a number of research out on vertical upward sediment laden jets. A


group of them aims at studying the dynamics of volcanic ash and understanding
its deposition around the volcano. Carey et al. (1988) carried out experiments on
vertical sediment laden upward plumes with low and high sediment concentra-
tions. A plume with low sediment concentration behaves as a pure plume with
reduced buoyancy flux. Re-entrainment of sediment back to the plume results in
increased particle flux along the plume.
Sparks et al. (1991) studied the the deposition from the radial surface spread-
ing current of dilute vertical upward sediment-laden plumes. Falling sediment
particles were observed to be drawn towards to the plume and may be re-
entrained. A simple model was developed to predict the sedimentation from
the spreading current with sediment mass conservation. The model compared
well with their experimental data.
Zarrebini and Cardoso (2000) studied the sediment deposition from the spread-
ing current of a vertical buoyant jet. They developed a similar model with Sparks
et al. (1991), but accounting for particle re-entrainment. The model explained
their experimental data well. Sedimentation from the margin of the plume is nev-
ertheless not accounted for. A similar study on vertical sediment-laden upward
pure jets is reported in Cardoso and Zarrebini (2002). Enhanced entrainment at
the surface is observed for pure jets.
Ernst et al. (1996) studied the particle fall out from the margin of vertical
buoyant jet and plume with experimental measurement and theoretical mod-
elling. For small particles and strong jets with large buoyancy, sediment re-
entrainment was observed and the effect had to be empirically accounted for in
their model. Similar study on the deposition of vertical upward sediment-laden
momentum jets was carried out by Neves and Fernando (1995) and empirical
equations were developed using dimensional analysis to predict the deposition
around the jet nozzle.
All the previous studies are limited in that they only focus on one or two of the
important mechanisms involved in the sedimentation process, including fall out
from spreading current, fall out from jet margin and reentrainment. The complex
sediment dynamics of vertical sediment laden buoyant jets cannot be properly
reflected in a integral jet model without introducing assumption and empiricism.
In particular, the buoyant jet induced flow in the ambient has hitherto not been
properly accounted for in a general way. A model which can incorporate all these
important mechanisms is necessary for a satisfactory prediction of sediment-laden
buoyant jets.

15
2.3.3 Horizontal sediment-laden buoyant jets

Bleninger et al. (2002) studied the bottom deposition of particle-laden horizontal


momentum round jets in stagnant ambient by experiments. They developed
a model to estimate the 2D bottom deposition pattern based on dimensional
analysis and an analytical equation to solve the particle motion. Particles are
assumed to fall out from the jet when the terminal settling velocity is greater
than the jet entrainment velocity. Particles are tracked using the time-averaged
longitudinal and radial velocity superimposed on the particle settling velocity. A
Gaussian transverse deposition profile is assumed. However, their experimental
data showed huge scatter as their sediment supply method could not guarantee
constant sediment concentration. The model prediction did not compare well
with experiment data as particle-turbulence interaction is not modelled.
For a buoyant jet with the jet trajectory bending up, the settling mechanism
is much more complicated. Two-modes of settling mechanism from horizontal
buoyant jets has been identified in previous studies. A narrow, concentrated
deposition region is found near the source, representing the fall out from the
jet edge. As the jet bends over, sediment is carried upwards to the surface
spreading current. Sediment settling out from spreading current forms a large
and near circular deposition region.
McLarnon and Davies (1999) studied the dynamics of a dilute particle-laden
plane buoyant jet with finite volume experimentally (as a starting buoyant jet
with a thermal-like cap and plume like tail). Dimensional analysis was carried
out to relate the bottom deposition characteristics to jet and sediment properties.
Lane-Serff and Moran (2005) developed an integral model to describe the sed-
imentation from a turbulent, buoyant round jet into a stationary ambient. Sedi-
ment is assumed to fall out from the jet where the normal component of settling
velocity to the jet boundary exceed the entrainment velocity. For the spreading
current fall out, Sparks et al. (1991)s model is used. A uniform concentration
of sediment is assumed in the jet, which is not supported by experimental obser-
vation. Jet-induced external flow field is not considered. Only six experiments
were carried out to validate the model.
Based on the idea of Lane-Serff and Moran (2005), Li (2006) and Lee et
al. (2012) further developed a two-layered sediment jet integral model using
JETLAG (Lee and Cheung, 1990) to account for the heterogeneity of sediment
concentration and the sediment mass transfer between the layers. It has consid-
ered the external flow, though crudely, to track the sediment trajectory outside
the jet. Several empirical factors are required to model the particle-turbulence
interactions.
Li (2006) and Lee (2010) carried out extensive experiments on horizontal
sediment-laden momentum and buoyant jets in stagnant water respectively. The
experimental set-up features an hourglass sediment supply system for effective
control of source sediment concentration, which is a significant improvement
to previous investigations using mechanical stirrer for maintaining sediment in
suspension. Lee (2010) developed a particle imaging method to measure the

16
cross-sectional sediment concentration in horizontal sediment-laden momentum
jets. The method is extensively compared with suction sampling measurement
and cross-checked against the bottom deposition profile. The heterogeneity of
sediment concentration in the upper and lower layers of the jet is experimentally
supported.
In addition, Lee (2010) also developed a 3D stochastic particle tracking model
to predict the sediment bottom deposition and concentration in horizontal mo-
mentum jets. Nevertheless, the model cannot compare well with experiments.
Cuthbertson et al. (2008) carried out experimental and CFD modelling on
the deposition from sediment-laden plane buoyant jets in stagnant ambient.
They found that jet trajectories are unaffected by source sediment concentra-
tion < 0.1% by volume. Two modes of deposition were identified: near source
deposition indicative of particle fall-out from the buoyant jet margins, and depo-
sition from spreading gravity current. Using dimensional analysis, the deposition
distributions can be described using the jet specific buoyancy flux, source den-
simetric Froude number and sediment settling velocity. Eulerian CFD model
prediction showed that the source configuration and lateral boundary confine-
ment can significant affect the trajectory of settling particles.
Cuthbertson and Davies (2008) studied experimentally the deposition from
particle-laden, round, buoyant jets in stationary and coflowing ambient. They
utilized photographic techniques to measure the bottom deposition. Similar to
their study for plane buoyant jet in stagnant waters, the two main sediment fall-
out mechanisms, the fall-out from jet margin and from the spreading current,
are identified. Dimensional analysis is used to quantitatively characterize the
deposition distributions.

2.3.4 Sediment-laden jet applications in ocean dumping


One of the engineering application of sediment-laden jet is ocean dumping of
dredge spoil or sludge. Sediment is dumped either instantaneously through an
ocean barge or continuously through pipes. For instantaneous dumping, the
sediment cluster acts as an instantaneous source of negative buoyancy, form-
ing a thermal as it travels down the water column. For continuous dumping,
the sediment flow acts as a high-concentrated sediment-laden jet with negative
buoyancy.
Koh and Chang (1973) developed a mathematical model to predict the trans-
port and deposition for instantaneous and continuous dumping. The model ac-
counts for three sediment transport phases: (1) the convective jet or thermal
phase influenced by initial momentum and negative buoyancy, (2) the collapse
phase as sediment falls out from the jet or thermal, and (3) the far field advective-
diffusion phase.
Many experimental and theoretical studies have been carried out on instan-
taneous thermal-like dumping (e.g. Li, 1997; Buhler and Papantoniou, 2001;
Bush et al. 2003; Gensheimer et al. 2012). Three phases of sediment dynamics
are identified: (1) the particle cloud accelerates under initial momentum and

17
buoyancy, transforming it into a well defined sediment thermal; (2) the thermal
with self-similar internal vortex circulation entrains ambient fluid and grows in
size; (3) as the vortex circulation and turbulence weaken, sediment segregates
from the thermal and enters the swarm phase, falling under their stillwater set-
tling velocity with weak dispersion. CFD and integral models were developed to
predict the sediment thermal dynamics.
Unlike the present study of sediment-laden jets in dilute concentration, these
sediment thermal studies are carried out at much higher sediment concentration;
thus sediment-induced buoyancy is not negligible.

2.3.5 Prediction of sediment impact of ocean outfalls


Sewage carries sediment of varies sizes. Sediment removal ability depends on
sewage treatment level. For preliminary screening, solids > 200m size are re-
moved while smaller solids are discharged through submarine outfalls. These fine,
organic-rich sediment, may deposit close to the outfall and negatively impact the
benthic ecology. Models have been developed for prediction and assessment of
the fate and transport of sediment from submarine outfalls. Farley (1990) de-
veloped a two-layer model for predicting the sediment deposition from an ocean
outfall in stratified environment. The model incorporatef tidal transport and
mixing, fine sediment coagulation and sediment-nutrient dynamics. The model
was validated against measured field data near an ocean outfall in California,
USA. Cromey et al. (1998) developed a model for predicting sediment deposi-
tion from sewage discharge and its biological impact. The model used stochastic
particle tracking to simulate sediment transport and settling. It incorporated
models of sediment resuspension, organic carbon chemistry and benthic biology.
The model is applied for environmental assessment of a British marine outfall.
These models account only for far field sediment dynamics.
As a microscopic view, Zhang and Li (2006) researched on the coagulation and
sedimentation of fine particulate in sewage. They hypothesized that higher sed-
iment concentration at the jet nozzle increases the flocculation of fine sediment,
thus increases the settling velocity and sediment settles close to the discharge.
Nevertheless, such hypothesis has not been supported by any experimental or
field observation of sediment-laden jets.

2.3.6 Summary of previous studies of sediment-laden jets


Previous studies of sediment-laden jets is carefully reviewed. The following can
be summarised:

1. Extensive experimental studies on dilute sediment-laden buoyant jets have


been carried out, covering a wide range of sediment and jet flow conditions.
The general features of sediment-laden jets is fairly well-known, but a more
fundamental understanding of sediment fall-out and deposition mechanisms
is lacking.

18
2. Analytical and numerical integral models have been developed to explain
the experimental data. Nevertheless there is hitherto no general model to
satisfactorily predict the bottom deposition of a sediment-laden buoyant
jet.

2.4 The equation of motion of sediment particle


Physical mechanisms governing particle motion has to be understood for pre-
dicting sediment dynamics in turbulent flows. The equation of motion consists
of the forces acting on a particle, including gravitational/buoyancy, fluid drag,
fluid acceleration due to local pressure gradient, added mass and Basset history
force (see Chapter 3).

2.4.1 Sediment settling velocity


The settling velocity of a sediment particle in stillwater ws is normally deter-
mined from the balance of drag and gravity. The drag force depends on particle
diameter, particle and fluid densities and fluid kinematic viscosity. An expression
relating drag coefficient and the particle Reynolds number Rep = ws d/ can be
obtained (see Chapter 3).
For natural sediment or irregular particles, the drag force deviates from the
spherical drag law and depends on the shape of the particle. Empirical equations
are often used to correlate the settling velocity with the particle size derived from
median sieved diameter (e.g. Hallermeier, 1981; Soulsby, 1997).
Stillwater settling velocity is widely used in environmental flow modelling,
such as sediment transport modelling in rivers. Nevertheless, particle-turbulence
interaction on settling velocity received little attention.
For a group of particles settling in quiescent water, Bulher and Papantoniou
(2001) studied the sediment dynamics when they enter the final stage of settling
- the swarm/cluster stage. It was observed that the particles descends in their
stillwater settling velocity and the surrounding fluid is motionless. Though the
sediment concentration is low enough, the width of the swarm still increases due
to the diffusion effect formed by the wake of the falling particles and their own
random lateral motion during the descent.

2.4.2 Basset history force


The Basset history force (Basset, 1888) accounts for the viscous effect due to
the delay in boundary layer development. Whenever there is a change in the
relative velocity between particle and fluid, the fluid around the sphere exerts an
additional unsteady force by viscosity. Since the evaluation of the Basset force
involves the integration of particle-fluid relative velocity accelerations through-
out its time history, it is computationally demanding with additional memory
storage.

19
Brush et al. (1964) derived analytical solution for the unsteady motion of
particle settling under linear drag with the Basset force. Numerical solution is
required for the non-linear drag range. Several studies aimed at developing more
efficient/more accurate method in computing the Basset force have also been
reported (e.g. Michaelides, 1992; Alexander, 2004; Bombardelli et al. 2008)
While most previous studies using Lagrangian particle tracking method ig-
nore the Basset force term, some studies showed that it can be important for
certain conditions. Michaelides (1997) pointed out that the necessity to include
the transient force terms (Basset, added mass, fluid acceleration) depends on the
quantity of interest of prediction. For an integrated quantity, like particle disper-
sion or time-averaged particle velocity, the transient force terms make negligible
contributions because their values change signs frequently and their integrated
effect becomes very small. On the other hand, the Basset force may be an im-
portant portion of the instantaneous total force on the particle. Therefore the
conclusions of previous studies on Basset force have to be viewed with caution.
For example, Mei et al. (1991) found that the Basset force has little effect on
particle diffusivity; but Liang and Michaelides (1992) showed the Basset force is
about 20% of the instantaneous total force on the particle when the fluid to par-
ticle density ratio is comparable to one and when the particle size is very small.
Bombardelli et al. (2008) showed that the Basset force considerably lengthens
the trajectory of a saltating sand grain.
The previous studies on Basset force are limited in several aspects. The linear
equation of motion is assumed (Rep << 1). The application of Basset term to
larger particles (Rep > 1) is problematic. These previous theoretical studies are
carried out for simple flow conditions only. Most importantly, these findings
are only based on theoretical model computations, without any experimental
support. The importance of Basset force has hitherto not been addressed for
turbulent sediment-laden jet flows.

2.5 Particle settling in turbulent flows


Many previous studies showed that particles settle more slowly in a turbulent
flow. Ho (1964) measured the settling velocities of spheres in homogeneous verti-
cally oscillating fluid. The apparent settling velocity can be reduced to as much
as 40% of the settling velocity in stillwater, when the fluid acceleration is about
10g, where g = gravitational acceleration.
Murray (1970) and Nielsen (1993) carried out particle settling experiments
on oscillating grid turbulence. They showed that the apparent settling velocity is
significantly modulated by turbulence; when the ratio of turbulent fluctuations
to terminal settling velocity is less than 3-4, a reduction of settling velocity up
to 30% is observed. Recent studies by Doroodchi et al. (2008) and Zhou and
Cheng (2009) have also shown a maximum settling velocity reduction of 25%
under grid-turbulence.
Hwang (1985) derived analytical solutions to explain the experimental results

20
of Ho (1964) and Murray (1970). Mei (1991) carried out stochastic numerical
simulation to study the effect of turbulence on particle settling velocity and
concluded that settling velocity reduction is due to non-linear drag. Stout et
al. (1995) carried out numerical prediction using the equation of motion to in-
vestigate the effect of non-linear drag on settling velocity in oscillating flows and
homogeneous turbulence. It was found that settling velocity reduction depends
on the particle Reynolds number Rep , ratio of turbulent fluctuation to settling
velocity (/ws ) and the frequency of oscillation.
In turbulent flows, the eddies are not simple sinusoidal unidirectional fluctua-
tions but in the form of rotating eddies rapidly varying spatially and temporally.
It is not necessary that reduction of settling velocity in turbulent eddies is entirely
caused by the non-linear drag effect. Stommel (1949) first found that small set-
tling particles (algae) follow closed trajectories in idealized circulation cells and
thereby remain suspended. Tooby et al. (1977) observed experimentally that
a sphere moves in closed circular trajectory in a forced vortex. Nielsen (1984)
showed analytically that some circular vortex flows (e.g. forced vortex, Rankine
vortex) can result in the trapping of sediment particles (Fig. 2.2).
Turbulence arises from shear instabilities at large Reynolds numbers and dis-
sipated by fluid viscosity. Turbulent eddies are random and three-dimensional
with a wide spectrum of time and length scales. Similar to the idealized vor-
tices, the 3D eddies would induce sediment trapping. Nielsen (1992) postulated
that a sediment particle may be trapped in a turbulent eddy during the eddy
formation and released until the eddy dissipates. Nielsen developed an velocity
autocorrelation function to describe this effect, but his assumption has not been
substantiated experimentally or examined in a turbulent shear flow such as a
sediment jet.

21
Figure 2.2: Trapping of sediment particle in a forced vortex adapted from Nielsen
(1984).

2.6 Mathematical modelling of two-phase flows


Eulerian and Lagrangian models have been applied in the study of particle trans-
port in turbulent flows. The characteristics and advantages of both modelling
techniques are reviewed and discussed.

2.6.1 Eulerian modelling


Eulerian two-phase modelling solves the continuity and momentum equations
for both fluid and solid phases. Momentum transfer between phases is incorpo-
rated into the momentum equations as source/sink terms. For a 3D two-phase
application, at least 10 partial differential equations (PDEs) have to be solved
simultaneously for two-phase flows (1 continuity and 3 momentum equations for
each phase, 2 for typical k- turbulence model). This poses high demand to
computation power, and most importantly, the turbulent interaction between
liquid and solid phases has not been understood thoroughly. Such modelling ap-
proach is only typically used in high-concentration two-phase flows in industrial
application.
For dilute sediment-laden flows in the environment, the solid phase does not
significantly affect the fluid phase. The solid phase transport is described by
mass conservation equation only. This approach is called a mixture model. A
settling flux is included in the vertical direction using a drift velocity (usually
taken as the stillwater settling velocity). The modelling of sediment transport
in the environment usually adopts this approach (e.g. Ji, 2009). Modelling the

22
sediment-induced buoyancy can be included in the equation of state and as a
damping function in the turbulent closure model. The approach is reported in
some studies on cohesive sediment transport (Toorman, 2002; Winterwerp and
van Kessel, 2003). Computation using the mixture model is more efficient as
only 7 PDEs have to be solved, but the use of stillwater settling velocity is
questionable.

2.6.2 Lagrangian modelling


The Lagrangian modeling approach solves the trajectories of individual fluid par-
ticles in a stochastic manner. In his classical analysis, Taylor (1921) showed that
the turbulent velocity fluctuation of fluid particles in successive time steps pro-
cess certain autocorrelation. The autocorrelation function can be approximately
described by a exponential decay function
 
hu (t)u (t + t)i t
R(t) = 2
= exp (2.1)
TL
where TL is a characteristic Lagrangian time scale of turbulence
Z
TL = R( )d (2.2)
0

Using the known velocity autocorrelation function, the velocity fluctuation of


individual fluid particles can be modelled stochastically. The turbulent diffusivity
in the Eulerian modeling can be related to the RMS turbulent velocity fluctuation
and Lagrangian time scale by
D = 2 TL . (2.3)

For sediment movement in turbulence, the situation is complicated by particle


inertia and its natural settling tendency. Csanady (1963) is among the first ones
to address this problem systematically. The settling of a particle under gravity
results in the crossing trajectory effect; as the particle does not follow the
turbulent eddy motion, the sediment particle autocorrelations are different from
that of fluid particles. By ignoring the particle inertia effect, Csanady derived
an autocorrelation function for a particle falling in atmospheric turbulence:
v
u 2 2
t u w T
R(t) = exp t1 + s 2 L (2.4)
TL Lz

where ws is the settling velocity; Lz is the turbulence length scale in the z


direction. It can be seen that when the particle is neutrally buoyant (ws = 0), the
correlation reverts to that of Taylors correlation. Whenr ws > 0, the correlation
ws2 TL2
is smaller than Taylors due to the positive term of 1+ L2z
. As explained
by Csanady (1963), this means owing to the finite mean settling velocity, the
heavier particle continuously changes its fluid-particle neighbourhood. The heavy
particle falls out from the eddy and therefore will lose its velocity correlation more
rapidly than a fluid particle which changes its turbulent velocity only owing to the

23
eddy decay. Csanadys autocorrelation function is supported by experimental
measurement of the particle Lagrangian velocity in grid turbulence by Snyder
and Lumley (1971).
The analysis of Csanady has certain limitations despite its pioneering contri-
bution. Firstly, it does not include particle inertia effect and assumes particles
respond to fluid motion instantaneously. This assumption may be plausible for
small-sized particles falling in air as the solid-air density ratio is usually large
(> 1000). The inertia effect may be significant for large particles in water. Sec-
ondly, the autocorrelation function cannot model the loitering and trapping effect
of particles in turbulence.
Besides based on the autocorrelation function, another type of stochastic
modelling is the eddy interaction time model proposed by Gosman and Ioan-
nides (1983). The eddy interaction time TI is defined as the minimum of
turbulent time scale TE and the residence time of a particle in an eddy TR ,
representing the cross trajectory effect,

TI = min(TR , TE ) (2.5)

where
LE LE
TE = =q (2.6)
2k/3
LE
TR = (2.7)
|up uf |
k 3/2
LE = C (2.8)

where C is a calibration constant; LE is the length scale of turbulence and
|up uf | is the particle-fluid relative velocity magnitude. The fluctuation ve-
locity is selected randomly from a Gaussian distribution with variance equal to
the RMS turbulent velocity . The particle is assumed to be in the eddy for
the duration of the interaction time, and then a new eddy with lifetime and
scale corresponding to the local turbulence field is established and a new fluctu-
ation velocity is randomly generated. This model has been integrated with CFD
models using the k turbulence closure.

2.7 Summary
In this chapter, previous studies of sediment-laden buoyant jets have been re-
viewed. For a single-phase turbulent buoyant jet, its trajectory and mean veloc-
ity can be robustly predicted using an integral jet model. The jet RMS turbulent
velocity is self-similar and can be predicted from the mean flow properties. The
jet-induced external flow can be modelled by a distribution of point sinks of
strength equal to the entrainment per unit length along the jet trajectory. The
surface spreading current formed by an impinging jet is extensively studied and
characterized. The understanding and predictability of these three important

24
flow regimes paves the way to the general modelling of sediment-laden buoyant
jets.
The governing equation of particle motion is reviewed. The importance of
Basset history force has hitherto not been studied on sediment-laden jets. Sedi-
ment motion in turbulence is discussed. The trapping and loitering of a sediment
particle in turbulent eddies has not been examined in sediment-laden jets. Eu-
lerian and Lagrangian modelling approaches of two-phase flows are discussed.
Previous sediment-laden jet studies have been reviewed. They contributed
some insights on the physics of sediment-laden jets. Various models have been
developed for sediment-laden jets. Nonetheless, the models are specifically de-
veloped for particular types of sediment-laden jets. There exists no satisfactory
general model for predicting the mixing and bottom deposition of an arbitrarily
inclined sediment-laden buoyant jet.

25
26
Chapter 3

Theory: Particle Tracking Model

3.1 Introduction
In this chapter the development of a three-dimensional stochastic particle track-
ing model is presented. The core of the particle tracking model is (1) the equation
of motion of sediment and (2) the Lagrangian autocorrelation function which
models the turbulent fluid velocity fluctuations along the sediment particle tra-
jectory. The chapter begins with the equation of motion of sediment particle
and the physical meaning of each force term. Secondly, the development of the
stochastic model on particle-turbulence interaction is discussed in detail. A de-
tailed validation of the model is presented, showing that the model is capable
of predicting sediment motion in various flow conditions. The influence of the
computationally demanding Basset force term on the settling velocity reduction
is discussed.

3.2 The governing equation of particle motion


3.2.1 Components of the equation of motion
The equation of motion of a spherical particle in an unsteady, non-uniform fluid
field is given as according to Maxey and Riley (1983):
dup
p Vp = (f p )Vp g (3.1)
dt
1
f CD Ap |up uf | (up uf )
2 !
duf
+f Vp
dt f
!
dup duf
f CM Vp
dt dt f
Z t d(up uf )
3
d2 f d d
2 0 t

27
where

up =
particle velocity = (up , vp , wp )
uf =
fluid velocity = (uf , vf , wf )
Vp volume of particle = d3 /6
=
Ap projected area of particle = d2 /4
=
p =
particle density
f =
fluid density
g (0, 0, g) = gravitational acceleration, g = 9.81 m/s2
=
CD =
drag coefficient = f (Rep )
CM =
added mass coefficient = 0.5
d =
particle diameter
|up uf |d
Rep = particle Reynolds number =

= kinematic viscosity of fluid = /f
= dynamic viscosity of fluid

The left hand side of the equation denotes the acceleration of the sphere and
the right hand side of the equation represents the forces acting on the spherical
particle. The terms are (i) body force (gravity/buoyancy), (ii) drag force, (iii)
fluid acceleration due to local pressure gradient, (iv) added mass force and (v)
Basset history force. For a solid particle in steady motion, only the gravitational
and drag forces are present. The fluid acceleration, added mass and Basset
history forces appear for unsteady particle motion. Except for the gravitational
acceleration which is self-explanatory, here the physical meaning of each of the
force terms are illustrated.

Drag

The drag
1
f CD Ap |up uf | (up uf )
2
is the force exerted by the fluid on the particle due to skin friction and pressure
difference between the upwind and downwind faces of the sphere. It depends on
the projected area of the particle, the relative velocity between particle and fluid,
and the fluid density and viscosity. The drag coefficient for spherical particle CD
is a function of the particle Reynolds number Rep based on the relative velocity
of particle and fluid.
d|up uf |
Rep = (3.2)

where is the kinematic viscosity of the fluid. For small Rep < 0.4, the drag
force, thus CD can be analytically determined using potential flow theory as
24/Rep . For large Rep > 1000, CD is equal to a constant value of about 0.45. In
between the relation of CD and Rep is obtained by experimental measurement.

28
A mathematical expression is fitted for modelling purposes. The CD Rep curve
used here is (Fig. 3.1)
24   0.42
CD = 1 + 0.15Re0.687
p + (3.3)
Rep 1 + 42500Re1.16
p

This relation is adopted by a number of studies as it covers a wide range of


particle Reynolds numbers. The expression corresponds to the limiting values in
the Stokes (laminar) drag range for small Rep (< 1) and the constant of 0.42 in
large Rep . For the particles used in the present study, their Rep is in the order
of 1-20.
10000
CD Stokes

1000 Clift and Gauvin


(1970)

100

10

Rep
0.1
0.01 0.1 1 10 100 1000 10000 100000

Figure 3.1: Drag coefficient-particle Reynolds number relationship by Clift and


Gauvin (1970).

Fluid acceleration

The pressure gradient !


duf
f Vp
dt f
due to unsteady fluid motion (e.g. spatially varied flow, turbulence) act as force
to the particles. The local pressure gradient gives rise to a force in the direction
of the pressure gradient. It is important
  to note that the derivative represents
duf
the fluid local pressure gradient dt as if the particle is absent, i.e.
f
!
duf uf
= + (uf )uf (3.4)
dt f
t
 
duf
It is not the same as dt
, which is the pressure gradient along the path of a
p
sediment particle, i.e.
!
duf uf
= + (up )uf (3.5)
dt p
t

29
Only the former is meaningful to the physical interpretation of the equation of
motion. The two are only equivalent in case of non-spatially varying flow, i.e.
! !
duf duf uf
= = , (3.6)
dt f
dt p
t

like a homogeneously oscillating fluid (e.g. Ho, 1964). In the case of sediment-
duf
laden turbulent jets, the term dt is prescribed using the mean flow only,
f
as the stochastic model does not involves the modeling of spatial turbulence
structures. The term can be analytically calculated with the analytical mean
flow solutions at any position. However, it is shown that the fluid acceleration
term is negligible in predicting the deposition of sediment-laden jets.

Added mass

The added mass force !


dup duf
f CM Vp
dt dt f

is an additional force that the surrounding fluid has to be displaced at the expense
of work done by the particle. The value of the added mass coefficient of a spherical
particle is theoretically determined as 0.5 (Lamb, 1932), which means that the
mass of fluid needed to be displaced equal to half of the mass of fluid occupied
by the sphere.

Basset history force

The Basset history force


Z t p d(u u )
3 2 f

d f d d
2 0 t
accounts for the viscous effects as the delay in boundary layer development as
the relative velocity changes with time. It represents the viscous shear force
acting on the particle as there exists a velocity gradient between the moving
particle and the stagnant ambient. It is called history force as it depends on
the acceleration history up to the present time. Since the evaluation of the
Basset force involves the integration of the accelerations throughout its time his-
tory, it is computationally demanding with additional storage needed for storing
the previous accelerations. It has to be noted that the Basset term is derived
based on laminar flow around the sphere thus it is applicable only to Stokes flow
(Rep < 1). The application of Basset term to larger particle with higher Rep is
problematic. In this study Rep > 1 for all the particles used in the turbulent
jet experiment. Nevertheless the importance of the Basset force is still evaluated
using the original expression, as the Rep are less than 20, for which the flow at
the vicinity around the sphere are still in laminar state. In the later chapters,
importance of the Basset force term will be further discussed.

30
3.2.2 Solution of the equation of motion
The equation is to be solved by numerical integration together with the particle
position.
dxp
up = (3.7)
dt
In sediment jet applications, the particle position provides the mean flow velocity
and the turbulent properties (RMS turbulent velocity and dissipation rate). The
numerical method used in the present study to solve the equation of motion and
the treatment of the Basset term, is described on Appendix A.1 and A.2.
The unknown in the equation of motion is the particle velocity up which is
dependent on the particle properties (density, diameter), and the fluid velocity.
In unsteady turbulent flow, the fluid velocity is particularly important as it gov-
erns the motion of the particle. At each time step, the fluid velocity has to be
known for the prediction of the particle velocity. In a stochastic random walk
model, the turbulent fluctuations has to be generated based on the autocorrela-
tion of turbulent fluid velocities along the particle trajectory, which is derived in
the next section.

3.3 The autocorrelation function of a sediment


particle in turbulence
3.3.1 Taylors autocorrelation function
The theoretical consideration starts from Taylor (1921)s classical analysis on
turbulent diffusion. As turbulence is in the form of coherent eddies, a particle
moves along with an eddy until the eddy dissipates and is taken up by another
one. Therefore the motion of particles is not completely random but certain
autocorrelation exists for the particle velocity at two consecutive times. The
Lagrangian correlation function of the turbulent fluctuating velocity u in a ho-
mogeneous turbulent field with zero mean flow is represented by

hu (t)u (t + t)i
R(t) = (3.8)
2
where t is the current time; t is the time difference and is the root-mean-
square (RMS) turbulent velocity fluctuation; hi denotes the ensemble average.
According to Taylor, in homogeneous turbulence, the Lagrangian autocorrelation
function R of the velocity of a neutrally buoyant particle following the fluid flow
can be described by an exponential function,
 
t
R(t) = exp (3.9)
TL

where TL is the Lagrangian integral time scale for turbulence. For small time
difference t 0, the particle velocity is completely correlated with itself (R =

31
1). At large time difference t , the particle losses its memory at the
previous times with correlation drops to zero.
For a foreign particle like sediment, the velocity autocorrelation also exists
but the drift velocity and particle inertia complicate the form of the function,
thought it is still partly following the fluid motion.

3.3.2 Derivation of the equivalent Eulerian particle veloc-


ity autocorrelation function
The derivation of the equivalent Eulerian particle velocity autocorrelation func-
tion is based on the work of Nielsen (1992). ConsideringD the Eensemble mean of
the absolute value of the time derivative of the velocity du
dt

, by definition
* +
du
h|u|i = t =

t (3.10)
dt TL

Correspondingly the autocorrelation function Eq. 3.9 may be written as


!
h|u|i
R = exp (3.11)

Consider the total velocity derivative,


du u u u u
= +u +v +w (3.12)
dt t x y z
Eq. 3.10 may be written as
* +
u u u u

h|u|i = +u +v + w t (3.13)
t x y z

To obtain the ensemble average of the Eulerian velocity increment, Nielsen


(1992) has made two assumptions:

1. the ensemble average depends on the velocity of the previous time step as
small turbulent velocity fluctuation correspond to small changes of velocity,
and

2. the cross-correlations
 between the velocity  components (u, v, w) and their
u u u u
partial derivatives t , u x , v y , w z are zero.

The first assumption is important to consider the structure of turbulence in trap-


ping the sediment particle by vortex motion, generating the so called loitering
effect as discussed in Nielsen (1992). The assumption is indeed a heuristic one
as different turbulent structures may affect the behaviour of sediment particles
in various ways. Experimental support is required to confirm the validity of the
assumption. Particle Imaging Velocimetry (PIV) measurement of the velocity of
turbulent jet flows has been carried out to show the validity of this assumption

32
in turbulent jet flows (see Chapter 4, Section 4.7.3). The second assumption can
also be substantiated using PIV velocity measurement (see Chapter 4, Section
4.7.4).
Thus the ensemble average is written as
v* + * + * + * +
u
u u u u u
h|u|i = tt ( )2 + u2i ( )2 + vi2 ( )2 + wi2 ( )2 (3.14)
t x y z

Applying the Eulerian equivalents to the ensemble values of the partial deriva-
tives, as like in Eq. 3.10 * +
u

= (3.15)
t TE
and * + * + * +
u u u

= = = , (3.16)
x y z LE
the ensemble average of the Eulerian velocity increment is derived as
v " #
u
t u
t1 + A ui 2 vi 2 wi 2
h|u|i = E ( ) + ( ) + ( ) (3.17)
TE

with the introduction of


TE
AE = . (3.18)
LE
where LE and TE are the Eulerian spatial and time scale of the turbulence
respectively. In turbulent jets and plumes, the Inserting back to Eq. 3.11, the
corresponding autocorrelation function is
v
u ! !2 !2
u 2
t u u i vi wi
Ri = exp t1 + AE + + . (3.19)
TE

The autocorrelation function depends on the velocity of the previous time


step, thus denoted by the subscript i. With the first assumption of Nielsen
(1992), the correlation is different from the one by Taylor which is a constant
relationship independent of the current velocity. This correlation can be viewed
as a local adaptation of Taylors correlation function in turbulence.

Estimation of turbulent length and time scales

The particle tracking model can be readily incorporated with the turbulent quan-
tities (k, LE and TE ) estimated from commonly used two-equation turbulence
closure models. Using the k model, LE and TE can be estimated according
to Launder and Spalding (1974), where

k 3/2
LE = C3/4 (3.20)

33
s
3 3/4 k
TE = C (3.21)
2
where k and are turbulence kinetic energy and its dissipation rates and C =
0.09 is the constant in the k model. The root-mean-square velocity fluctuation
can be related to the turbulence kinetic energy by
s
2
= k. (3.22)
3
With Eqs. 3.20 and 3.21,
TE
AE = =1 (3.23)
LE
In a jet with self-similar profiles of RMS turbulent fluctuations and turbulent
energy dissipation rates (see Chapter 5, Fig. 5.17), the turbulent length scales
LE can be shown as in the order of 0.1b, where b is the characteristic half-width
of a jet (Fig. 3.2).
If the first assumption is waived, the ensemble averages of the spatial deriva-
tives can be written as
* + * + * +
u u u 2

u = v = w = , (3.24)
x y z LE
With AE = 1, the autocorrelation function can be recovered as
 
2t
R(t) = exp (3.25)
TE
In this case, the expression is equivalent to Taylor original one with TL = 0.5TE .
Csanady (1973) suggest that TL = 0.25TE , while experimental data of Sato and
Yamamoto (1987) suggested that TL = 0.30.6TE for grid generated turbulence,
which is in good agreement with the derivation here.

3.3.3 The autocorrelation of sediment particle in turbu-


lence
In the present work, the Full Model solving the governing equation of motion
is proposed, using the velocity autocorrelation function derived for sediment
particles settling in turbulence. The Simplified Model is originally proposed
by Nielsen on modelling the trapping of sediment particles in grid turbulence; in
the present study it is extended to apply for sediment-laden jets.

Full Model

Sediment particles do not completely follow the fluid motion but with a relative
velocity due to the gravity or its own inertia (Fig. 3.3). In a turbulent flow, the
particle velocity up can be written as
up = uf + ur (3.26)
= uf + uf + ur

34
(a) Schematic illustration

0.4 0.4
/uc LE/bg

0.3 0.3

0.2 0.2

0.1 0.1

0 0
0 1 2 r/bg 3 0 1 2 r/bg 3
(b) Self-similar RMS turbulent velocity (c) Normalized Turbulent length scale
profile

Figure 3.2: Turbulent length scale LE variation of a round jet. (a) Schematic
illustration. (b) Self-similar RMS turbulent velocity profile estimated using
Eq. 5.7, Chapter 5. uc is the jet centerline mean velocity. (c) Turbulent length
scale normalized with Gaussian jet half-width bg = 0.114x, estimated using
Eqs. 5.7 and 5.8, Chapter 5.

35
where uf = uf + uf is the fluid velocity (mean+turbulent fluctuation) and ur is
the relative velocity between the particle and the fluid. Without loss of generality,
the time mean flow can be assumed as zero (uf 6= 0 in a turbulent jet flow) and
the particle velocity is
up = uf + ur (3.27)

Figure 3.3: Motion of sediment particle in fluid. xp is the particle position; up


is particle velocity; uf is fluid velocity at the location of particle; ur is relative
velocity.
On the solution of the equation of motion of particles, the turbulent fluctua-
tion of fluid has to be generated for each time step, based on the autocorrelation
of fluid velocity along the particle path. Thus it is important to obtain this
autocorrelation function.
By definition, the change of the fluid velocity along the particle path (take
the vertical component w as an example), denoted by the subscript p, is
! ! ! ! !
dwf wf wf wf wf
= + up + vp + wp (3.28)
dt p
t p
x p
y p
z p

Similar to Eq. 3.14, the ensemble average of the change of velocity is


v* + * + * + * +
u
u u u u u
h|u|i = tt ( )2 + u2p,i ( )2 + vp,i
2 2
( )2 + wp,i ( )2 (3.29)
t x y z
The mean properties of turbulence along the particle path should be the same
as the turbulent properties of the fluid. Therefore the mean values of the partial
derivatives are the same as Eqs. 3.15 and 3.16 (/TE , /LE ). Similar to Eq. 3.17,
the velocity change is
v " #
u
t u
t1 + A up,i 2 vp,i

w
2 + ( p,i )2
h|u|i = E ( ) + ( ) (3.30)
TE
The autocorrelation function can be expressed as
v " #
u
t u u v w
Ri = exp t1 + AE ( p,i )2 + ( p,i )2 + ( p,i )2 (3.31)
TE

36
Simplified Model

For a fluid particle following completely the fluid motion, up = uf = u , Eq. 3.19
is recovered. For a particle settling with constant velocity ws in the vertical
direction only, the relative velocity is

0
ur = 0

(3.32)
ws

With this assumption, the solid particles follows the fluid velocity completely
except in the vertical direction. The particle velocity is

uf

up = u + ur = vf , (3.33)
wf ws

the autocorrelation function can be shown similarly as


v " #
u
t u u v w ws 2
Ri = exp t1 + AE ( f,i )2 + ( f,i )2 + ( f,i ) (3.34)
TE

which is the same as derived in Nielsen (1992). Eq. 3.31 is applicable for the
Full model which the equation of motion is solved, while Eq. 3.34 is for the
simplified model with the assumption of constant relative velocity (Eq. 3.32),
with the assumption that the statistical properties of the particle velocity is the
same as the turbulent fluid velocity. It has to be noted that the simplified model
is based on the assumption of particles following the fluid velocity except in the
vertical direction which is the superposition of the fluid velocity and the particle
stillwater settling velocity. Nevertheless the loitering and trapping of particles in
turbulent eddies are generated by the autocorrelation function itself (see Section
3.3.5).
Interestingly, if the first assumption of Nielsen is waived, like Eq. 3.24, the
autocorrelation would becomes independent of the instantaneous velocity fluctu-
ations and can be written as
v
u
t u w2 T 2
R(t) = exp t4 + s 2 E (3.35)
TE LE

which is indeed similar to the form given by Csanady (1963), derived for particles
falling in atmospheric turbulence, with the inertia of particles ignored.
v
u
t u w2 T 2
R(t) = exp t1 + s 2 L (3.36)
TL Lz

Eqs. 3.31 and 3.34 appears to be applicable to modeling both loitering effect as
suggested by Nielsen and the cross-trajectory effect as suggested by Csanady.

37
3.3.4 Generation of turbulent velocity fluctuations
The turbulent diffusion process is often viewed similar to a molecular diffusion
process (see e.g. Pope, 2000). With the autocorrelation function, the turbulence
fluctuation can be generated by
q
uf ,i+1 = Ri uf ,i + (1 Ri2 ) (3.37)

where the subscript i denotes the current time step and i + 1 denotes the next
step. denotes randomly generated numbers (in x, y, z-directions) following a
Gaussian distribution with zero mean and unit variance. With the instantaneous
particle velocity determined, the particle position can be found by numerically
integrating the velocity with respect to time (Eq. 3.7). In the Full Model, the
particle position is integrated with the equation of motion using a predictor-
corrector method (see Appendix A.1). The particle position (Eq. 3.7) in the
Simplified Model is integrated using a predictor-corrector approach.
1
xi+1 = xi + (ui+1 + ui )t. (3.38)
2

3.3.5 Physical interpretation of the autocorrelation func-


tion
Eq. 3.34 can be viewed as a heuristic local correlation that mimics the vortex
trapping of sediment particles demonstrated by Nielsen (1992). From Eq. 3.34,
there is an asymmetry between updrafts (w > 0) and downdrafts (w < 0). On
an updraft w and ws tend to cancel, so the argument of the exponential is closer
to 0 and correlation is greater (the exponential function is closer to 1). A particle
would tend to retain any upward motion for longer, whereas for a downdraft the
motion becomes decorrelated more quickly. Thus on average particles stay in the
upward moving flow longer than in the downward moving flow, resulting in the
reduction of settling velocity.
Extensive numerical experiments have confirmed the characteristic feature of
Eq. 3.34. Fig. 3.4 shows a time series of one realization of turbulent velocity
fluctuation of a settling particle in homogeneous turbulence (in the z direction
only, u , v = 0) with the following parameters: stillwater settling velocity ws =
0.01m/s; RMS turbulent fluctuation = 0.01m/s; turbulence length scale LE =
0.005m; turbulence time scale TE = LE / = 0.5s. Computation time step is
taken as 0.001s. The velocity fluctuation generated by Nielsens autocorrelation
function dependent on instantaneous velocity is compared to that generated by
a constant autocorrelation function
s
t w2
R(t) = exp 2 + 2s
TE

which is obtained by substituting the instantaneous values by the RMS turbulent


fluctuation ; it is similar to the autocorrelation by Csanady (1963) (Eq. 3.36).

38
The autocorrelation in the present model can be viewed as a local correlation
that varies with the instantaneous velocity (or turbulent fluctuation w ) in order
to capture the vortex trapping feature. This is heuristically reasonable in view
of the small time step as compared with the time scale of turbulence (i.e. the
instantaneous velocity is like a quasi-steady local velocity). When w and the
settling velocity ws cancel out, the autocorrelation reaches its maximum. The
turbulent velocity generated by Nielsens autocorrelation is slightly higher than
that by a constant autocorrelation. Fig. 3.5 shows the position of the settling
particle as integrated from the particle velocity wp = w ws . It can be seen
that the particle position using Nielsens function is always higher than that
by a constant autocorrelation and the particle is in general falls more slowly in
turbulence under the loitering effect. The reduction in settling velocity using
Nielsens autocorrelation is also supported by a number of realizations of the
particle position (Fig. 3.6).

0.05 1

0.04 0.998

0.03 0.996

0.02 0.994

Corr. Coeff (R)


0.01 0.992
w' (m/s)

0 0.99

-0.01 0.988

-0.02 w' (Nielsen R) 0.986


w' (constant R)
-0.03 Ws 0.984
R (Nielsen)
-0.04 R (Constant) 0.982

-0.05 0.98
0 0.5 1 1.5 2 2.5 t (s) 3

Figure 3.4: Time history of turbulent velocity (w ) and autocorrelation (R) of


a single realization, t = 0.001s. LE = 0.005m, = 0.01m/s, TE = 0.5s,
ws = 0.01m/s.

39
0
0 0.5 1 1.5 2 2.5 t (s) 3

-0.01

-0.02

z (Nielsen)
-0.03
z (Constant)
z (stillwater)
-0.04

z (m)
-0.05

Figure 3.5: Time history of particle position (z) of a single realization, t =


0.001s. LE = 0.005m, = 0.01m/s, TE = 0.5s, ws = 0.01m/s.

40
0
0 0.5 1 1.5 2 2.5 t (s) 3

-0.01

-0.02

-0.03
Constant settling

-0.04

z (m)
-0.05
(a) Nielsens autocorrelation
0
0 0.5 1 1.5 2 2.5 t (s) 3

-0.01

-0.02

-0.03

Constant settling
-0.04

z (m)
-0.05
(b) Constant autocorrelation

Figure 3.6: Ten realizations (black lines) of the time history of particle position
(z), t = 0.001s. LE = 0.005m, = 0.01m/s, TE = 0.5s, ws = 0.01m/s. Red
thick line is constant settling velocity.

41
3.4 Model validation
The particle tracking model is validated against four different tests: (1) accel-
erated motion of particles in stagnant water, (2) motion of particles in homoge-
neously oscillating fluid, (3) motion of particles in homogeneous grid turbulence,
and (4) vertically downward sediment-laden jets.

3.4.1 Accelerated motion of a particle in stagnant fluid


The accelerated motion of a sediment particle in stagnant water is the first ba-
sic validation test to the particle tracking model, as analytical solution can be
24
derived for the case of linear (Stokes) drag law (CD = Re p
). Ignoring the Basset
history term, the time variation of the particle velocity can be derived with zero
initial velocity of zero
!
wp (t) 18t
= 1 exp 2 (3.39)
ws d (s + CM )

where ws is the stillwater settling velocity

(s 1)gd2
ws = (3.40)
18

The analytical solution of an accelerating sphere in the Stokes range with the
Basset force can be found in Brush et al. (1964) in a closed form solution.

wp (t) c2 + h2 h
= 1+ exp(h2 t) exp(c2 t) sin(2cht )erfc(c t) (3.41)
ws h
s
Z h i
t
2 exp(y 2 t) cos[2c(h y)t ]dy
0
where

9
c =
2d(s + CM )
3 q
h = [8(s + CM ) 9]
2d(s + CM )
!
1 h
= tan
c

and erfc(x) is the complementary error function.


2 Z
erfc(x) = exp(t2 )dt

Fig. 3.7 shows the comparison of analytical and numerical solutions of the
motion of a 50m diameter sphere falling in water with the parameters p =
2500 kg/m3 , f = 1000 kg/m3 , = 106 m2 /s. The stillwater settling velocity

42
ws is 2.044 mm/s. For this small particle, the time to reach its terminal settling
velocity is in the order of 0.001s, thus a time-step of 2 104 s is about one fifth
of the transient period and can be considered as coarse time step. The numerical
solution lies close to the analytical solution, for both time steps of 1 104 s and
2 104 s. This ensures the correctness of the numerical implementation.
Significant difference can be observed for the motion of the particle with
and without the Basset force. Without the Basset term, the particle reaches its
stillwater settling velocity at about 0.002s. With the Basset term solution, the
particle accelerates at a much slower rate, only reaching 95% of the stillwater
one at 0.1s. The Basset force is important in the accelerating motion of particles.
It decays much more slowly than the added mass force due to the slow infinity-
approaching nature of the Basset integrand. This contributes an additional drag
to the particle and considerably lengthens the time for particle to approach its
still water settling velocity. However, this example only shows the effect of Basset
force for a particle moving uni-directionally. The overall effect of the Basset force
on the motion, dispersion and deposition of sediment particles in oscillating flow
and turbulence has to be demonstrated by further numerical experiments.
As the analytical solution is only strictly applicable to the Stokes drag law,
validation of the model in the non-linear drag range requires experimental data.
Brush et al. (1964) reported a number of experiments which the velocities of
settling particles are measured by dropping spheres in liquids. In the results only
the density ratio s and the particle Reynolds number Rep is given thus the particle
diameter is inferred from the given paramaters. A total of six experiments are
presented. In the numerical model the time step is 0.001s for all the cases.
Fig. 3.8a shows the comparison of the numerical model with the experimental
data with a normalized time without Basset force. The comparison is considered
satisfactory as the general trend of the time history of the particle velocity is
captured, despite the model result shows a shorter time for reaching the steady
state for some of the cases. A simulation with the Basset force (Fig. 3.8b) shows
that the comparison is slightly better at the beginning of the particle motion.
However, at the later stage of particle motion, the Basset force seems to delay
the steady motion of the particle, thus the particle velocity is lower than the
measurement. The Basset force may not be strictly applicable in the case of
high particle Reynolds number as the viscosity effect is no longer dominating the
flow around the sphere.

3.4.2 Motion of particles in homogeneously vertically os-


cillating fluid
Ho (1964) studied the motion of falling spherical particles in a homogeneously
vertically oscillating fluid. The experiments were conducted in a cylinder of 45.7
cm long and 8.9 cm inner diameter. The cylinder was filled up with fluids with
different density and viscosity, and oscillated vertically with different amplitudes
and frequencies. Spheres were released into the cylinder during the experiments
and the particle motion was recorded on a camera to determine its velocity. Three

43
1.2
w/ws

0.8

0.6

0.4
Analytical w/o Basset
Analytical w/ Basset
0.2
Numerical dt = 0.0001s
Numerical dt = 0.0002s
0
0.0001 0.001 t (s) 0.01

Figure 3.7: Comparison between analytical (Eqs. 3.39 and 3.41) and numerical
solutions of particle falling in stagnant fluid, d = 50m, p = 2500 kg/m3 , f =
1000 kg/m3 , ws = 2.044mm/s

different fluids and five sizes of particles of various specific gravity were used to
give a combination of 11 conditions of particle Reynolds number Rep based on
the stillwater settling velocity and two density ratio conditions (s = p /f = 7.8
and 2.5).
In the numerical model, a particle is released at rest in the oscillating fluid
for which the time-dependent motion is described by the equation:

wf (t) = wf 0 cos(t) (3.42)


= 2/T

where wf 0 is the velocity amplitude of cylinder oscillation; T is the period of


oscillation. The motion of the particle is computed for a number of oscillation
cycles (at least 50) until the velocity variation is less than 0.1% different from the
previous cycle. Typical time step is chosen as 1/50 of the period of an oscillation
cycle. The average settling velocity is computed by averaging the velocities in
an oscillation cycle. Four of the Rep conditions: Rep = 1.1, 6, 28 and 230 are
simulated and compared with the experimental data as these particle Reynolds
number is similar to those used in the present study. For each case of Rep , there
were two oscillation period conditions same as the experiments (normalized as
/d2 , where = 2/T ); and by varying the velocity amplitude, a number of
experiments were carried out for each frequency condition. Table 3.1 shows the
eight cases used for validation of the particle tracking model.

44
w/ws w/o Basset
1

0.8

0.6

0.4
Re p = 540, d = 2028m m
0.2 Rep = 153, d = 1026m m
Re p = 28, d = 440mm
0
0.001 0.01 0.1 T = t/d
2 1
(a) Without Basset force

w/ws w/ Basset
1

0.8

0.6

0.4
R e p = 540, d = 2028mm
0.2 R e p = 153, d = 1026mm
R e p = 28, d = 440mm
0
0.001 0.01 0.1 T = t/d
2 1
(b) With Basset force

Figure 3.8: Comparison of numerical model prediction and experimental data of


particle falling in stagnant fluid (Brush et al. , 1964). s = p /f = 2.5.

45
Fig. 3.9 shows the comparison between the model predicted ratio of apparent
settling velocity to terminal settling velocity (wsa /ws ), as a dependent variable
of the ratio of oscillation acceleration to the gravitational acceleration (wf 0 /g)
for the four cases of Rep . The comparison between model and experiment is very
good. For all the cases, a general trend can be observed that, as the oscillation
amplitude (thus acceleration) increases, the reduction in wsa becomes more sig-
nificant. Also, for the same acceleration, the settling velocity reduction increases
with Rep . For example when wf 0 is about 7 time of the gravity acceleration, the
settling velocity reduction is about 30% for Rep = 1.1. This increases to about
40% for Rep = 6 and even 60% when Rep = 280. Conversely, the dependence of
velocity reduction on the parameter /d2 is relatively small with only noticable
difference at the larger Rep .
As shown, with the Basset force included, the apparent settling velocity is
only slightly different. The difference is larger for experiments with the largest
Rep of 230, about 7-8% of ws . For Rep = 1.1, 6 and 28, which is the range
of particle Reynolds number in the present sediment-laden jet experiments, the
difference is mostly less than 5%.
The settling velocity reduction can be attributed to the particle inertia during
the oscillating motion of the fluid. Fig. 3.10 shows the time history of particle
and fluid velocity, and the forces acting on the particle. Due to particle inertia,
the particle velocity always lags behind the fluid velocity. This results in the
variation of the drag force as the relative velocity between particle and fluid
varies. Since the drag force does not vary linearly with the relative velocity
for large particles (Rep > 0.4), the enhanced drag force results in the overall
reduction of the settling velocity. On the other hand, the added mass only
contributes an insignificant portion of the forces.

Table 3.1: Experiments of Ho (1964) of spheres settling in homogeneously verti-


cally oscillating fluid.

ReT diameter solid fluid s viscosity Settling Oscillation /d2 ws


m density density p /f 106 velocity Period
kg/m3 kg/m3 m2 /s cm/s s
230 1201 2650 1000 2.65 0.929 17.79 0.20 0.021
230 1201 2650 1000 2.65 0.929 17.79 0.16 0.016
28 3175 9358 1206 7.76 35.77 31.54 0.19 0.110
28 3175 9358 1206 7.76 35.77 31.54 0.15 0.085
6 1587 9527 1206 7.90 35.77 13.52 0.19 0.430
6 1587 9527 1206 7.90 35.77 13.52 0.15 0.330
1.1 3175 9801 1263 7.76 250.8 8.69 0.19 0.740
1.1 3175 9801 1263 7.76 250.8 8.69 0.15 0.590

46
wsa /ws Rep = 1
1

0.8
2
/d
0.6
0.74-w/ Basset
0.59-w/ Basset
0.4
0.74-w/o Basset
0.59-w/o Basset
0.2
0.74-exp
0.59-exp wf0 /g
0
0 1 2 3 4 5 6 7

(a) Rep = 1
wsa/ws
1 Rep = 6

0.8
2
/d
0.6 0.43-w/ Basset
0.33-w/ Basset
0.4 0.43-w/o Basset
0.33-w/o Basset
0.2 0.43-exp
0.33-exp wf0 /g
0
0 1 2 3 4 5 6 7

(b) Rep = 6
wsa/ws
1 Rep = 28

0.8

2
0.6 /d
0.11-w/ Basset
0.4 0.085-w/ Basset
0.11-w/o Basset
0.085-w/o Basset
0.2
0.11-exp
0.085-exp wf0/g
0
0 1 2 3 4 5 6 7

(c) Rep = 28
wsa/ws
1 Rep = 230

0.8

0.6
2
/d
0.021-w/ Basset
0.4 0.016-w/ Basset
0.021-w/o Basset
0.2 0.016-w/o Basset
0.021-exp
0.016-exp wf0/g
0
0 1 2 3 4 5 6 7

(d) Rep = 230

Figure 3.9: Comparison of the numerical model and the experimental results of
Ho (1964) of spheres settling in homogeneously vertically oscillating fluid.

47
m/s
2.5
2
1.5
1
0.5
wa, -0.08
0
-0.5 0ws, -0.19 0.2 0.4 0.6 0.8 1
-1
Particle
-1.5 Fluid
-2
-2.5

(a) Fluid and particle velocities


Force
8
Gravity
6
Fluid
4 Drag
2 Added mass

0
0 0.2 0.4 0.6 0.8 1
-2

-4

-6

-8

(b) Forces acting on the particle, normalized by gravity

Figure 3.10: The particle and fluid velocities, and forces acting on the particle.
Rep = 230, s = 2.65, T = 0.2s, wf 0 = 2m/s.

48
3.4.3 Motion of particles in homogeneous grid turbulence
The test of the model of the auto-correlation relationship is on a random turbu-
lent field. Oscillating grid turbulence is commonly used to study the basic char-
acteristics of turbulence for the calibration of turbulence closure models. The
turbulence generated by an or a pair of oscillating grid(s) resembles the charac-
teristics of homogeneous isotropic turbulence (Cheng and Law, 2001). Fig. 3.11 is
an schematic representation of dropping a particle in turbulence. Murray (1970)
and Nielsen (1993) reported experiments of particle settling under grid-generated
turbulence and measured the particle velocity. Zhou and Cheng (2009) carried
out particle imaging velocimetry measurement on sediment settling in grid tur-
bulence. Numerical model simulation is carried out and comparison is made to
Murray (1970) and Zhou and Cheng (2009)s experiments.

N = 1000 releases
w s = 0.02 m/s

H = 1m
T = STi / N
1m
wsa = H / T

Figure 3.11: Schematic representation of a hypothetical experiment of dropping


sediment particles in homogeneous turbulence.

Laboratory Experiments

Murray (1970)
The experiment was carried out in a 50cm wide channel with 40cm water
depth. Three grids with 5 cm grid size, separated 30cm apart are oscillated in
phase to generate the turbulence. The oscillation stroke is 40cm with speeds
from 16 to 45 cm/s. Particles with nominal diameter 2mm with various densities
are used to obtain different settling velocities in water (Table 3.2). A neutrally
buoyant particle with 2mm diameter are used to measure the RMS fluid veloc-
ity fluctuation. High-speed images were taken to evaluate the particle velocity.
Particle settling velocity reduction up to 60% is observed.

Zhou and Cheng (2009)

49
Table 3.2: Particles used in grid turbulence experiments of Murray (1970). =
106 m2 /s

diameter particle density fluid density Settling velocity Rep


d p f ws = ws d/
mm kg/m3 kg/m3 cm/s
2.0 1010 1000 1.0 20
2.0 1027 1000 2.0 40
2.0 1049 1000 3.0 60
2.0 1075 1000 4.0 80

Zhou and Cheng (2009) studied the motion of a single particle in grid turbu-
lence. The experiments are carried out in a tank of 50 50 80 cm height. A
single grid 5cm size, located at the mid-depth, is oscillated vertically at a stroke
of 10.4 cm and frequencies of 2 and 3 Hz (grid velocity amplitude of 65 and 98
cm/s). Spherical particles of 6.35 and 7.94mm with density p = 1050 kg/m3
(ws = 8.85 and 10.29 cm/s) are used. In addition, cylindrical particles (p =
1077 kg/m3 ) of 2.4 mm diameter and 3 mm in length, ws = 2-6 cm/s are used
(Table 3.3). The particle velocity is measured by tracking the sediment particle
position in subsequent images. Particle Imaging Velocimetry (PIV) is used for
measuring the fluid velocity by seeding it with neutrally buoyant 20 m particles.
Their /ws ratio is modest, ranging from 0.05 to 0.4 and the result shows a mean
reduction of about 10% of the stillwater settling velocity.

Table 3.3: Particles used in grid turbulence experiments of Zhou and Cheng
(2009)

diameter particle density fluid density Settling velocity Rep


d p f ws = ws d/
mm kg/m3 kg/m3 cm/s
6.35 1050 1000 8.85 562
7.94 1050 1000 10.29 817
2.4 (cylindrical) 1077 1000 2-6 48-144

Numerical Predictions

Numerical experiments are carried out to investigate the effect of homogeneous


turbulence on the settling velocity. In each hypothetical experiment, N = 1000
particles are allowed to settle for a distance of H = 1m in a field of homogeneous
turbulence and zero mean flow. The turbulence properties are specified with
the RMS turbulent velocity fluctuation and the turbulence length scale LE .
Various ratio of /ws are specified. LE is taken as 0.1 of the grid size, both
LE = 5mm for of Murray (1970) and Zhou and Cheng (2009)s experiments. TE
is computed from Eq. 3.23 with AE = 1 specified. The time for falling for 1m for

50
each particle Ti is recorded and the apparent settling velocity wsa is calculated
from the mean settling time.

H
wsa = 1 PN (3.43)
N i=1 Ti

Fig. 3.12 shows the cumulative probability distribution of the apparent set-
tling velocity of numerical experiments using glass particle of 180m and plastic
particles of 716m under different to ws ratios. With increasing turbulence
level, the mean apparent settling velocity decreases but the variance in apparent
settling velocity increases. Turbulence results in both dispersion of particle and
a reduction in the mean settling rate. With the same /ws (e.g. 1 and 5), plastic
particles have a greater reduction in settling velocity under turbulence, despite
their similar stillwater settling velocity with glass particles.
Fig. 3.13 shows the comparison of apparent settling velocities as a function of
/ws with the experimental data of Murray (1970). The model prediction follows
broadly the trend of the experimental data, despite the considerable scattering.
As turbulent level increases, the apparent settling velocity decreases. For the four
particles, their velocity reduction is about the same with slightly larger reduction
for the heaviest particle. Fig. 3.14 shows the model comparison with the data
of Zhou and Cheng (2009) for which the ws ratio is less than 0.4. Very little
settling velocity reduction is predicted by the model, corresponding well to the
measured apparent settling velocity. The considerable scattering also illustrates
the uncertainties in these oscillating grid turbulence experiments. Nevertheless
the model follows broadly the trend of settling velocity reduction measured in
independent experiments.
The particles used in the sediment jet experiments in this study are smaller
in size than those used in previous grid-turbulence experiments. Fig. 3.15 shows
the settling velocity in different to ws ratios for the glass particles G115, G180
and G215 and the plastic IP3 particle. LE = 0.005m is assumed for all the
cases. Reduction of settling velocity can be observed as /ws increases. For
/ws < 1 the trend of the four kinds of particles are about the same. For
/ws > 3, the largest sized glass particle G215 has the highest reduction in
settling velocity. For the smallest particle G115, the reduction is the smallest.
For the IP3 particles, the settling velocity reduces much more when compared
with the glass particles. This is attributed to the particle inertia of the larger
and lighter IP3 particles. The numerical prediction provides explanations to the
deposition profile difference in the sediment-laden jets using plastic particles (see
Chapter 5).
The difference between the full model and the simplified model is shown
in Fig. 3.15. While the predicted settling velocity reduction of both models lies
at the same curve for /ws < 1, the result deviate for /ws > 1, for which
the particle inertia has an effect on the settling velocity reduction. The pre-
dicted settling velocity reduction by simplified model shows a level off at about
80% beyond /ws > 1, while the prediction from the full model shows further
reduction.

51
Cumulative probability
1
G180
0.9 / ws
0.8 0.1
0.5
0.7
1
0.6 5

0.5

0.4

0.3

0.2

0.1
ws (m/s)
0
0.005 0.01 0.015 0.02 0.025
3
(a) Glass particle, d = 180m , p =2500 kg/m , ws = 1.95cm/s
Cumulative probability
1
IP3
0.9 / ws
0.8 0.1
0.5
0.7
1
0.6 5
0.5

0.4

0.3

0.2

0.1
ws (m/s)
0
0.005 0.01 0.015 0.02 0.025
(b) Plastic particle, d = 716m , p =1140 kg/m3 , ws = 2.02cm/s

Figure 3.12: Cumulative probability distribution of model predicted sediment


settling velocities under different /ws with AE = 1 and LE = 0.005m, (a) glass
particles, (b) plastic particles.

52
wsa/ws
1
0.9
0.8
0.7
0.6
1cm/s
0.5 2cm/s
0.4 3cm/s
4cm/s
0.3 1cm/s-exp
0.2 2cm/s-exp
3cm/s-exp
0.1 4cm/s-exp
0
/ws
0.1 1 10
Figure 3.13: The predicted apparent settling velocity compared with the exper-
imental data of settling particle in grid turbulence of Murray (1970).

wsa/ws

1.0

0.8

0.6

0.4

0.2 Zhou and Cheng (2009)

Model
0.0
0.01 0.1 /ws 1

Figure 3.14: The predicted apparent settling velocity compared with the exper-
imental data of settling particle in grid turbulence of Zhou and Cheng (2009)

53
1.2
wsa/ws w/o Basset
1

0.8

0.6
G180
0.4 G115
G215
0.2 IP3
Simp. Model
0
0.1 1 /ws 10
(a) Without Basset force
1.2
wsa /ws w/ Basset
1

0.8

0.6
G180
0.4 G115
G215
0.2 IP3
Simp. Model
0
0.1 1 /ws 10
(b) With Basset force

Figure 3.15: The predicted apparent settling velocity for the particles used in
this experiment, (a) without Basset force, (b) with Basset force.

54
Fig. 3.15(a) and (b) compare the effect of the Basset force on the overall
settling velocity reduction. The trend of all curves in both figures are very
similar, indicating that the Basset force has little contribution on the reduction
of settling velocity.

An expression for estimating the settling velocity reduction in CFD


models

Since the settling velocity reduction depends on the RMS turbulent velocity
which can be estimated with a computational fluid dynamics model using a k
turbulence closure, the predicted reduction is settling velocity from the particle
tracking model can be utilized to provide a functional estimate of the apparent
settling velocity in CFD calculation. The numerical results can be fitted in a
curve in the form of
wsa A
= (1 + A) (3.44)
ws B + (1 + B) exp(C/ws )

where A = 0.265, B = 0.459 and C = 1.285 are empirical constants. Fig. 3.16
shows the apparent settling velocities as a function of /ws . The experimen-
tal data of Murray (1970), Nielsen (1993) and Zhou and Cheng (2009) under
grid turbulence is shown together and lies quite close to the two curves, de-
spite considerable scatter. Since the RMS fluctuation is readily available using a
two-equation turbulence closure (such as k , assuming isotropic turbulence),
the reduction in settling velocity would be reasonably predicted using Eq. 3.44
and adopted in Eulerian modeling of sediment/particle transport (see Chapter
5, Section 5.4).

55
wsa/ws

1.0

0.8

0.6

0.4 Murray (1970)


Nielsen (1993)
0.2 Zhou and Cheng (2009)
Eq.14
0.0
0.01 0.1 1 /ws 10

Figure 3.16: The predicted apparent settling velocity, best-fitted with Eq. 3.44,
and compared with the experimental data of settling particle in grid turbulence
of Murray (1970), Nielsen (1993) and Zhou and Cheng (2009)

3.4.4 Vertically downward sediment-laden jets


A vertically downward jet is utilized to study the effect of turbulence on sed-
iment particles (Fig. 3.17). Particles are injected into the jet flow in the noz-
zle and carried by the flow downward, same as the direction of gravity, with
turbulent mixing with the ambient. As a simpler case of particle laden flows,
a number of studies has been carried out on vertical downward sediment laden
jets. Singamsetti (1966) carried out experiments using natural sand particles and
measured the radial particle concentrations in different jet longitudinal sections.
Parthasarathy and Faeth (1987) measured the particle velocity and particle flux.
The data from these experiments are used for validation of the particle tracking
model.

Experiments of Singamsetti (1966)

The vertical downward sediment-laden jet experiments were carried out in a


cylindrical tank of 1.83m (72) diameter and 1.22m (48) depth, with a jet
nozzle of D = 6.655mm (0.262) diameter. The jet velocity was 6.096 m/s (20
fps) for all experiment cases with different sediment diameter. Natural sands
with median diameter of 68, 115, 230, 320 and 460m were used. Sediments are
sampled using isokinetic suction sampling at various positions (all less than 50D)
along the jet centerline and radial positions for concentration analysis. Table 3.4
summarises the jet flow and sediment properties. It has to been noted that

56
Figure 3.17: A downward sediment laden jet.

the mean jet velocity is much higher than the sediment settling velocity at the
sampling locations.
Fig. 3.18 shows the model predicted sediment concentration for three cases
of sediment-laden jet and the solution of a pure jet. The sediment concentration
has close resemblance with that of a pure jet. Fig. 3.19 shows the comparison
of the centerline sediment concentration predicted by particle tracking model
and experimental measurement. For all the cases, the sediment concentration
decays with the -1 power law, same as the decay of tracer concentration. As
the fluid velocity is much greater than the settling velocity at the region x <
50D, particles exhibit tracer-like behaviour. Fig. 3.20 shows the comparison of
predicted and measured cross-sectional profiles of sediment concentration. The
model predicted profiles are slightly narrower, and the data compares better with
Gaussian profiles; nevertheless the general trend is well predicted. Singamsetti
(1966) suggested that the larger particles have a larger spread than the smaller
particles, but such phenomenon is also not observed in the model prediction.
The effect of the Basset force on sediment dispersion is compared for three of
the experimental cases (68, 230 and 460 m). Predictions (Fig. 3.19) show that
Basset force has very little contribution to the sediment concentration profiles.

57
0 0

0.15 0.05

0.6
0.5

0.3
0.05
0.1

0.05
0.0 0.2

20 20

0.1
0.3

0.05
0.2

0.2
0.01
0.1
0.01
0.01

0.01

40 40

0.05
z/D

z/D
0.1
0.05

0.05
0.1

0.1
60 60
0.05

0.01
0.1
0.01
0.01

0.01

0.05
80 80
0.1
0.05

0.05
0.05

100 100
-40 -20 0 20 40 -40 -20 0 20 40
x/D x/D

(a) 68m (b) 230m


0 0
0.5

0.05
0.4 0.4

0.05
0.1
0.05 0.1

0.3
0.2

20 20
0.01

0.1
0.05

0.1
0.2
0.01

40 40
0.01
0.05
0.01
0.1

0.05
z/D

z/D
0.05
0.1

0.01

0.1

60 60
0.05
0.01

0.01

80 80
0.05
0.01

0.05
0.01

100 100
-40 -20 0 20 40 -40 -20 0 20 40
x/D x/D

(c) 460m (d) Pure jet (analytical)

Figure 3.18: Predicted sediment concentration (C/C0 ) of vertical sediment laden


jet, cases of Singamsetti (1966), (a) 68m, (b) 230m, (c) 460m and (d) ana-
lytical pure jet tracer concentration.

58
10 10
C max /C 0 Cmax/C0
68m 115m

1 1

0.1 Model w/o Basset 0.1


Model w/ Basset
Free jet
Meas.
0.01 0.01
1 10 x/D 100 1 10 x/D 100

10 10
Cmax/C 0 Cmax/C0
230m 320m

1 1

0.1 0.1

0.01 0.01
1 10 x/D 100 1 10 x/D 100

10
C max/C 0
460m

0.1

0.01
1 10 x/D 100

Figure 3.19: Comparison of centerline sediment concentration for experimental


cases of Singamsetti (1966). Model prediction with Basset force is carried out
on cases of 68, 230 and 460 m, with very little difference with the one without
Basset force.

59
1.2 1.2
C/Cmax 68m C/Cmax 115m
1 1

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2
r/z
0 0
-0.2 -0.1 0 0.1 r/z 0.2
-0.2 -0.1 0 0.1 0.2

1.2 1.2
C/Cmax C/C max
230m 320m
1 1

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2
r/z r/z
0 0
-0.2 -0.1 0 0.1 0.2 -0.2 -0.1 0 0.1 0.2

1.2
C/Cmax 460m
1

0.8 Model w/o Basset


Model w/ Basset
0.6 Free jet
Meas.
0.4

0.2

0 r/z
-0.2 -0.1 0 0.1 0.2

Figure 3.20: Comparison of cross-sectional sediment concentration for experi-


mental cases of Singamsetti (1966). Model prediction with Basset force is car-
ried out on cases of 68, 230 and 460 m, with very little difference with the one
without Basset force.

60
Table 3.4: Experiments of Singamsetti (1966)

Case Jet Initial Sediment Settling Location where


velocity diameter velocity w s wf
a
m/s m cm/s in terms of D
1 6.096 68 0.374 10108
2 6.096 115 1.024 3689
3 6.096 230 3.199 1182
4 6.096 320 4.815 785
5 6.096 460 6.873 550
Jet diameter D = 6.655mm
a Estimated based on Soulsby (1997)s settling velocity formula for natural sediments.

Sediment density = 2650 kg/m3


Water density = 998.2 kg/m3 , water viscosity = 1.004 103 kg/m/s (20 C).

Experiments of Parthasarathy and Faeth (1987)

Two experiments of vertically downward sediment-laden jet were carried out in


a rectangular tank of 410 530 910 mm high with jet nozzle 5.08mm. Parti-
cle and liquid velocities are measured using phase-discriminating Laser Doppler
Anemometry (LDA) technique. Particle number flux in the longitudinal direc-
tion was measured. The two experiments have similar jet velocity but only differ
in the initial concentration (Table 3.5). 501m diameter spherical glass particles
were used in the experiments.

Table 3.5: Experiments of Parthasarathy and Faeth (1987)

Case Jet Initial Initial particle Sediment Settling Location where


velocity volume fraction diameter velocity ws wf
m/s (%) m cm/sa in terms of D
1 1.66 2.4 501 7.52 137
2 1.72 4.8 501 7.52 142
Jet diameter D = 5.08mm
a Estimated based on drag law for spherical particles.

Sediment density = 2450 kg/m3


water density = 997.1 kg/m3 , water viscosity = 0.8937 103 kg/m/s (25 C).

Model prediction and measured particle number flux are compared (Fig.
3.21). The predicted particle number flux is obtained by summing up of the
number of particles passes through an area. Theoretically the centerline tracer
mass flux decays in x2 manner as the centerline velocity and tracer concentra-
tion both decays in x1 manner. The measured centerline particle number flux is
higher than the x2 trend due to the particle inertia which inhibits the particles

61
from dispersion by turbulent eddies. The model predicted particle flux compares
very well with the data. The prediction shows that the mass flux remains nearly
constant until 10D, indicating a increased potential core length. This is sug-
gested also by Jiang and Law (2005) that the particle inertia results in lower
turbulent mixing. The comparison of the cross-sectional particle number flux
is shown in Fig. 3.22. The model compares very well with measurements and
highly resembles a Gaussian profile.

10
Nc/N 0

Model
0.1
Case I (P&F 1987)

Case II (P&F 1987)

-2
x
0.01
x/D
1 10 100

Figure 3.21: Comparison of predicted and measured centerline sediment mass


flux for experimental cases of Parthasarathy and Faeth (1987)

62
1.2 1.2
F/Fmax F/Fmax
z = 16D z = 24D
1 Model 1 Model

Case I (P&F
0.8 Case I (P&F 0.8 1987)
1987)
Case II (P&F
0.6 Case II (P&F 0.6 1987)
1987)

0.4 0.4

0.2 0.2

r/z r/z
0 0
-0.2 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 0.2 -0.2 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 0.2

1.2
F/Fmax
z = 40D
1 Model

Case I (P&F
0.8
1987)
Case II (P&F
0.6 1987)

0.4

0.2

r/z
0
-0.2 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 0.2

Figure 3.22: Comparison of predicted and measured cross-sectional sediment


sediment mass flux for experimental cases of Parthasarathy and Faeth (1987)

63
3.5 Summary
In this chapter, a particle tracking model is developed. The core of the model
is the governing equation of sediment particle motion and the turbulent veloc-
ity autocorrelation function for the fluid velocity along the particle path. The
autocorrelation function models the loitering and trapping of sediment particles
in turbulent eddies. Two particle tracking models are developed to predict the
movement of sediment particles in turbulent flows. The Full Model solves the
equation of motion numerically. The Simplified Model assumes the sediment
particle follows the fluid motion except at the direction of gravity. The particle
tracking model is validated with four basic tests. The model shows very good
comparison with previous experimental data. It is shown that the simplified
model works well for sand and glass particles with density 2500 kg/m3 , while
the full model is required for the modelling of particles with density close to wa-
ter. Model predictions have shown that the computationally demanding Basset
force term can be neglected in turbulent conditions.

64
Chapter 4

Experiments

The particle tracking model developed in the previous chapter is intended to


predict the transport and deposition of particulate matter in sediment-laden
buoyant jets. To test the applicability of the theory and study the gross fea-
tures of two-phase buoyant jets, experiments are carried out for validating the
developed model.
This chapter first describes the sediment-laden jet experiments carried out in
this study. The experiments were performed in the Croucher Laboratory of Envi-
ronmental Hydraulics at the University of Hong Kong. The experiments includes
horizontal particle momentum jets discharged in stagnant water with different
kind of particles injected. The experimental set-up is similar to that described
in Li (2006) and Lee (2010), except for the modifications to suit different types
of particles. Secondly, turbulent jet velocity measurement is carried out to give
experimental support to the loitering effect assumption in the autocorrelation
function developed in Chapter 3.

4.1 Experimental set-up

4.1.1 Experimental tank


A schematic diagram of the experimental set-up for horizontal particle-laden jets
is shown in Fig. 4.1. The set-up consists of a water tank of 1m 1m 0.5m high
made up of glass walls. The jet nozzle, fabricated with brass with an internal
diameter of 6mm is discharged horizontally. The jet nozzle is 15 cm long, located
at the mid-way of the side wall, and 0.25m above the bottom of the tank (0.15m
above the top of the sediment collection tray).

4.1.2 Jet discharge system


The jet discharge system consists of a storage tank, a constant head tank, a cal-
ibrated Tokyo Keiso rotameter to measure the discharge and a valve to regulate
the flow. The water is pump from the storage tank to the head tank with an

65
Figure 4.1: Experimental set up of horizontal sediment-laden jet.

1m x 1m x 0.5m Tank
hourglass
From particle feeder
Headtank Sediment
Sediment-laden jet

Rotameter Laser sheet


66

Argon Laser

bottom sediment
collection tray

High-speed camera

Data aquisition system


overflow to maintain constant pressure head. Fresh tap water is used for both
jet and ambient fluids in the horizontal momentum jet experiments. The source
and ambient temperatures are measured before each experiment to ensure that
their difference is less than 1 C to avoid unexpected buoyancy.

4.1.3 Sediment supply system


In many previous studies of sediment-laden jets, the sediment was usually pre-
mixed with the stored source fluid using mechanical mixers. This poses great
difficulty in providing constant initial sediment concentration for the jet, espe-
cially for heavier particles. Li (2006) developed a sediment supply method which
greatly alleviate this difficulty and improved the experimental accuracy. This
sediment supply system is adopted in this study. The sediment feeding system
consists of an hourglass and a vertical settling tube of 6mm internal diameter
(Fig. 4.2). Sediment is temporally stored in the hourglass and gradually fed into
the settling tube through a small opening. It can be shown experimentally (see
Fig. 4.4) and theoretically (Lee, 1981) that the sediment draining rate from the
hourglass is constant, dependent on the size of the opening only but independent
with the head of sediment above the opening. The particles released from the
hourglass fall under gravity into the settling tube and mix with the water, like
a buoyancy induced sediment thermal. It allows sufficient time for the solids to
be well-mixed before entering the jet flow stream. The solids are continuously
and constantly supplied from the hourglass, continuously travel down the set-
tling tube with mixing until it meet the jet flow, resulting a constant sediment
concentration discharged from the jet. Upstream of the sediment supply sys-
tem (about 20 cm), a mesh is installed inside the jet pipe to generate sufficient
turbulent mixing to prevent the segregation of particles inside the jet pipe.
To suit the properties of different kinds of particle, the design of the sediment
feeding system is different in their size, shape and installation location. Detailed
internal dimensions of the hourglasses are shown in Fig. 4.3. For glass particles
(Figs. 4.2a and 4.3a), the hourglass is located above the water surface of the tank,
thus particles travel for some distance in air under free fall before hitting the
water surface in the settling tube to form the sediment thermal. The hourglass
consists of a perspex tube and a brass assembly which the small opening is
located at. Three brass assemblies are manufactured with difference sized holes
for the release of glass particles in different diameter: 0.6mm for G115, 1mm for
G180 and 1.3mm for G215.
For plastic particles, the situation is more complicated due to the strong
surface tension between the particle and water. If the particles are released
above water it may only hit the water surface without sinking into the water, as
the particle surface tension force is stronger than gravity, clogging the settling
tube. A different design is adopted: the hourglass is partially-submerged below
the water surface of the tank (Fig. 4.2b). During each experiment, the particles
are pre-wetted and fed carefully into the hourglass sealed up with a stopper
which could equalize the pressure between the hourglass and the tank in the
neck. When the experiment starts, the stopper is removed and the particles are

67
released steadily through the neck into the settling tube, while the water in the
hourglass and settling tube is maintained in hydrostatic pressure.
The hourglasses, either installed submerged or above the water surface, are
shown to have a constant rate of discharging particles under initial testing exper-
iments. This is determined by measuring the time for a certain mass of particles
to be fully drained from the hourglass. Fig. 4.4 shows the time-mass discharge
relations for the submerged hourglasses for the plastic particles. The drainage
time-particle mass relationship is summarised in Table 4.1, with those measured
in the experiments of Lee (2010). The relationships provide estimation of the
time of an experiment with a known mass of particle, or vice versa, in order for
experimental planning. During each experiment, the experiment duration Texp
is still recorded to estimate actual initial jet sediment concentration C0 with the
equation
Mp
C0 = (4.1)
QTexp
where Mp is the mass of particle used; Q is the flow rate of the jet.

Table 4.1: Mass feeding rates for the hourglasses


Hourglass Type Opening diameter Sediment feeding rate
Particles (mm) (g/s)
a
IP3 Submerged 5.0 0.0337
a
MF Submerged 3.0 0.0111
G115b Surface 0.6 0.0206
G180b Surface 1.0 0.0534
G215b Surface 1.3 0.0928
a
This study.
b
Lee (2010).

4.1.4 Bottom sediment collection and measurement


The bottom sediment deposition rate is measured either one-dimensionally (trans-
versely lumped) or two-dimensionally. In some experiments, 1D profiles are de-
rived from the 2D measurement by summing the transverse deposition at the
same longitudinal location. For 1D profiles, a set of 26 rectangular removable
trays is used as sediment traps (Fig. 4.5a). The trays are fabricated with alu-
minum plates with sizes of 30cm length (y) 2.5cm height (z). For the first 20
trays laid close to the jet nozzle the width (x) is 2.5 cm, while for the last 6
trays, the width is 5 cm. This design of removable trays provides sufficient reso-
lution near the jet exit and the efficiency in sediment collection for 1D profiles.
The trays are placed inside one base tray made of perspex (83 cm length 32
cm width 7 cm height) and the top level of the removable trays are located
15cm below the jet nozzle.

68
(a) Hourglass for glass particles

(b) Hourglass for plastic particles

Figure 4.2: The two types of hourglass for feeding sediment in the experiment:
(a) for glass particles, (b) for plastic particles (IP3 and MF).

69
2 cm

6 cm
Perspex Perspex

1.7 cm
Brass
assembly

neck
Copper
(0.6mm, 1mm
Assembly
or 1.5mm dia.)
(a) Hourglass for glass particles

6 cm

7 cm

Perspex

3 cm
45
5 mm

neck
(5mm dia.)
(b) Hourglass for IP3 particles

3 cm

5.5 cm
Perspex

60 3 cm
5 mm

neck
(3mm dia.)
(c) Hourglass for MF particles

Figure 4.3: Detailed internal dimension of hourglasses

70
500
Time (s)
450
400
350 y = 29.638x
2
300 R = 0.9964

250
200
150
100
50 Hourglass for IP3 particles
0
0 5 10 15 20
Dry Mass (g)
(a) Hourglass for IP3 particles

800
Time (s)
700

600

500

400 y = 89.768x
2
R = 0.9891
300

200

100
Hourglass for MF particles
0
0 2 4 6 8 10
Dry Mass (g)
(b) Hourglass for MF particles

Figure 4.4: Time-mass discharge relationships of hourglasses for (a) IP3 particles
and (b) MF particles.

71
For measuring the 2D sediment deposition profiles, the base tray is made of
gluing a plastic screen board onto a plastic plate. The screen board consists of
square grids of size 1.43cm, with 58 grids in the longitudinal and 21 grids in the
transverse directions (x = 83cm, y = 30cm). The height of each grid is 2 cm.
Same as the 1D bottom tray, the top level of the 2D tray are located 15cm below
the jet nozzle (Fig.4.5b).
The base trays are coloured in black to increase the contrast between the
white coloured sediment and the tray for photographic documentation purpose
after each experiment. It also minimizes the reflection when a laser light sheet
is used for illumination.
After each experiment, the trays are removed from the ambient tank carefully
with minimum disturbance. Sediment particles fallen into each tray/grid are
collected with a syringe and transferred to small glass vials (4cm diameter
and 5cm height, Fig. 4.6). The mass of each glass vial is weighed before they
are used for experiments. In the 1D removable trays, the sediment mass at the
bottom tray is also collected as sediments may leak to the bottom tray through
the small gaps between the removable trays. The mass collected in the base tray
is usually below 2% of the total input mass. The glass vials with particles are
dried in an oven of 105 C for about 12 hours. After that they are taken out
and cooled to room temperature in a desiccator and weighed with an electronic
balance with 4 decimal-place accuracy to measure the sediment mass.

72
30 cm

6 trays @ 5 cm

20 trays @ 2.5 cm

Base tray

x
Removable
trays
y

Jet nozzle

(a) For measuring 1D deposition profiles

21 grids at 1.43 cm

35 grids @ 1.43 cm

Jet nozzle

(b) For measuring 2D deposition profiles

Figure 4.5: The two types of sediment collection trays for measuring: (a) 1D,
(b) 2D deposition profiles.

73
Figure 4.6: Glass vials for collecting sediment for weighing (diameter = 4cm;
height = 5cm).

4.2 Particle properties


The present study aims at studying the behaviour of sediment particles with dif-
ferent density and size in turbulent jet flows. Three kinds of particles are used,
including spherical glass particles, spherical polymethylmethacrylate resin parti-
cles (IP3) and irregular melamine formaldehyde particles (MF). The properties
of particles are summarised in Table 4.2.

4.2.1 Glass
The glass particles are spherical with density of 2500 kg/m3 , appears as transpar-
ent to white colour. Three sizes are used for experiments, with median diameters
115, 180 and 215m. The particles have a narrow size range for each class. Since
a number of experiments, including 1D bottom deposition and sediment concen-
tration measurement has been carried out using these glass particles as reported
in Lee (2010), only a few of them are repeated for checking of consistency and they
are not reported in this thesis. Nevertheless, six experiments (two experiments
for each particle size) are carried out to determine the 2D bottom deposition
profiles.

4.2.2 Polymethylmethacrylate, IP3


The plastic particles of polymethylmethacrylate resin (denoted as IP3) are spher-
ical with density of 1140 kg/m3 , appears in translucent white colour. The particle
has a median diameter of 716m. The particles have a narrow size range. Pre-
wetting is required and a submerged hourglass is used for storing them in the
experiments. Six experiments are carried out for 1D deposition measurement
and three for 2D measurement.

74
4.2.3 Melamine formaldehyde, MF
The melamine formaldehyde (MF) are irregular shaped particles with density
of 1500 kg/m3 . The particles are multi-coloured. They are sieved with British
Standard sieves into the range of 300-425m, then washed with tap water to
remove soluble impurities and oven dried before using them for experiment. Like
IP3 particles, MF particle requires wetting by water and a submerged hourglass
is used for storing them in the experiments. A total of six experiments are
carried out using MF particles, with three for 1D measurement and three for 2D
measurement.

4.2.4 Measurement of settling velocity and determination


of particle equivalent diameter
For the irregular MF particles, its sieve size range (300-425m) is considered too
wide for use in the particle tracking model using the equation of motion. Since
the size distribution is unknown, the mean or median of this size range, 363m,
is not suitable to be taken as a characteristic size. An equivalent diameter
deq for MF, and also the irregular natural sand particles particles in Li (2006)
is defined for the application in the equation of motion. deq is defined using the
expression for estimating the terminal settling velocity of spherical particles,
3CD
deq = ws2 (4.2)
4g(s 1)

where s = p /f is the specific gravity; g = 9.81m/s2 ; ws is the measured settling


velocity; CD = f (deq , ws , ) is the drag coefficient for spherical particles (Eq. 3.3)
and f and are the density and kinematic viscosity of water respectively which
is dependent on the temperature of water. CD and deq can be determined itera-
tively.
The particle settling velocities for MF particles are determined by measuring
the time the particle traveled downward for a known distance. The measurement
is carried out with a perspex cylindrical settling column with diameter of 7cm
and height of 65cm, filled with tap water. The vertical dimension is marked by
a tape ruler. A small cluster of particles is introduced into the water column
and the time for the middle of the cluster to travel for a distance by 20cm at the
mid-level of the column is measured. The experiment is repeated for 50 times to
obtain the mean settling velocity and its standard deviation. Water temperature
in the settling column is measured to determine the water density and viscosity
for calculation of the equivalent diameter. For the spherical particles (Glass
and IP3), the actual median diameter is used.

75
Table 4.2: Properties of particles in the present experiments.
Particles Shape Density Mean measured Standard Equivalent Size
(kg/m3 ) settling velocity deviation diameter rangea
(cm/s) (cm/s) (m ) (m )
G115 Spherical 2500 2.64 0.15 115 100-132
G180 Spherical 2500 1.83 0.13 180 156-209
G215 Spherical 2500 1.03 0.09 215 191-243
IP3 Spherical 1140 2.06 0.20 716 617-832
MF Irregular 1500 2.20 0.17 347 N/A
a Lee (2010).

4.3 Flow visualization


Flow visualization is carried out for a number of experiments to observe the
dynamics of sediment particle and the fluid. The flow visualization system com-
prises of a laser light sheet for illuminating the flow and the particles, with a
high-speed camera for capturing particle motion (Fig. 4.1). The light source is
an Argon-ion laser (Coherent Innova 90) with maximum power of 7 Watt with
wavelengths 514-488nm. The laser is turned on to warm up for one to two hours
before experiments and kept for 6.5 W for every experiment. A laser sheet is
formed as a vertical plane along the jet centerline by directing the laser beam
through a series of mirrors and a cylindrical lens. The position of the laser sheet
is well adjusted prior to the experiment for covering the whole area of interest.
Typical laser sheet thickness is about 2mm.
Upon illumination, the sediment particles appears as brilliant blue-green
colour, same as the light of the laser. For some of the experiments, fluores-
cent dye Rhodamine 6G (appears as bright yellowish colour when illuminated)
is added to the source fluid for the observation of the fluid phase turbulent flow.
Images of the experiments was captured by a 10 bit high speed CMOS camera
(PCO 1200 hs) with 2 Gigabyte memory and resolution 1280x1024 pixel. For
some of the preliminary experiments a Sony video camera with 24 frames per
second is used to capture video recordings of the experiment. The experiments
are carried out with all the lights switched off to eliminate background light and
maximize the lighting effect of the laser.

4.4 Fluid density and viscosity calculation


Fluid density and viscosity is required for the full particle tracking model and the
calculation of terminal settling velocity for the simplified model. In the present
experiments, the water density is determined from the measured temperature
using the well-known UNESCO formula (Millero and Poison, 1981). Before each
experiment, the temperature in both the supply tank and the water tanks is
measured by a FLUKE 54 II contact thermometer to ensure both source and

76
ambient temperature difference are less than 1 C to prevent unexpected buoy-
ancy. With the salinity of tap water close to zero, the density of freshwater can
be calculated by

(S, T ) = T + s (4.3)
2
T = 999.842594 + 6.793952 10 T
9.095290 103 T 2 + 1.001685 104 T 3 (4.4)
6 4 9 5
1.120083 10 T + 6.536332 10 T
s = AS + BS 1.5 + 4.8314 104 S 2
A = 0.824493 4.0899 103 T + 7.6438 105 T 2
8.2467 107 T 3 + 5.3875 109 T 4
B = 5.72466 103 + 1.0277 104 T 1.6546 106 T 2

where (S, T ) is the water density (kg/m3 ), T and S are the temperature ( C)
and salinity (ppt) respectively (S = 0). Viscosities of freshwater under differ-
ent temperature can be obtained from reference tables (Smithsonian Institution,
1954).

4.5 Summary of experiments


The experiments carried out in this study are summarized in Tables 4.3(glass),
4.4(IP3) and 4.5(MF). The run number designations listed on the tables follows
the systems as this: The first part designate the particle type - G for spherical
glass particles, with the later number as the particle size in m, IP3 for plastic
polymethylmethacrylate resin particles and MF for melamine formaldehyde par-
ticles. J stands for jets and the two-digit number is the jet flow rate in L/h. A
suffix of -2D indicates that 2D deposition profiles are measured.
The experiments cover a wide range of flow rates from 30L/h (u0 = 0.29m/s)
to 90L/h (u0 = 0.88m/s) in comparison with previous experimental studies
which has a much narrower range of jet flow velocity and Reynolds numbers
(e.g. Bleninger et al. 2002: Re = 700 1400, Lane-Serff and Moran, 2005:
Re = 3300 3600). The experimental flows has Reynolds numbers over 1500 to
ensure turbulent flow condition for all the experiments. Due to limited tank size
and avoiding jet interaction of the back wall, the jet Reynolds numbers Re of the
experiments are mainly around 2000-6000. Although the Res are much smaller
than the prototype (Re 200, 000), jets are turbulent for Re > 2000 and the jet
properties are not dependent on Re (e.g. Fischer et al. 1979).
The experimental procedures are:

1. Before each experiment, a certain mass of sediment is weighed and care-


fully transferred to the hourglass with a stopper preventing the release of
sediment.

2. The bottom tray is placed in the tank and aligned precisely with the jet.

77
3. The temperature in the source and ambient water are measured prior to
the experiment to ensure that the temperature difference is less than 1 C
to avoid unexpected buoyancy.
4. The experiments are initialized with the ambient kept stagnant, which is
achieved by waiting for at least 1 hour to damp out any disturbance.
5. The jet flow is first introduced to the tank without sediment added, for 2-3
minutes to stabilize the flow field.
6. Sediment is then introduced through the hourglass and this marks the start
of experiment. Time measurement (for estimating the sediment concentra-
tion) starts as the sediment starts to come out from the jet and ends as the
sediment ceases to appear. Each experiment typically lasts for 5 minutes.
7. Video or high-speed camera recordings are carried out for selected experi-
ments with room light switched off.
8. Photos are taken for the bottom deposition profile immediately after the
experiment, for documentation record and future cross checking against
deposition data.
9. The bottom collection trays are then taken out of water carefully. For the
1D removable trays, they can be taken out one by one, but for the 2D
tray, it has to be taken out of water with great care as a whole to avoid
resuspension of sediment during the movement.
10. The sediment in each tray/grid is transferred to the glass vials carefully
using a syringe of 2mm nozzle.
11. The glass vials with sediment are oven dried for 12 hours, then taken out
and cooled to room temperature and measured for the mass.

For each experiment, the mass recovery percentage RM is checked to ensure


that there is no mis-measurement of the mass.
total mass collected after experiment
RM = 100% (4.5)
Total mass input
The mass recovery of the present experiments are mostly over 90%. The mass loss
can be attributed to the loss during the bottom tray removal, and the transfer of
sediment from bottom tray to the glass vials. The 1D sediment deposition rate
(g/m/s) in a tray i is given as
Mi
Fs,i = (4.6)
Texp x
where Mi is the measured sediment mass in the tray and x is the longitudinal
size of the tray (2.5cm or 5cm). The 2D sediment deposition rate (g/m2 /s) in a
grid is given as
Mi
Fs,i = , (4.7)
Texp xy
where x and y are the bottom grid sizes.

78
Table 4.3: Horizontal momentum jet experiments using glass particles, measuring
2D bottom depositions.
Run u0 Re C0 a ws lm RM
No. (m/s) (g/L) (kg/m3 ) (cm/s) (m) (%)
G215J90-2D 0.8842 5603 3.82 997.7 2.65 0.177 97.2
G215J60-2D 0.5895 3735 5.14 997.7 2.60 0.121 95.8
G180J80-2D 0.7860 4980 4.78 997.7 2.04 0.205 93.8
G180J50-2D 0.4912 3142 7.64 997.6 2.05 0.127 97.4
G115J60-2D 0.5895 3805 1.04 997.5 1.00 0.315 94.1
G115J40-2D 0.3930 2490 1.77 997.7 0.98 0.213 93.8

Jet diameter = 6mm


u0 = Jet initial velocity
Re = Jet Reynolds number
C0 = Jet initial sediment concentration
a = Water density in the ambient
ws = Stillwater settling velocity predicted from drag law for spherical particles
1/2
lm = Momentum-settling length scale = M0 /ws = (/4)1/2 u0 D/ws
RM = Mass recovery ratio

Table 4.4: Horizontal momentum jet experiments using IP3 particles, measuring
1D and 2D bottom depositions.
Run u0 Re C0 a ws lm RM
No. (m/s) (g/L) (kg/m3 ) (cm/s) (m) (%)
1D deposition profile measurement
IP3J80 0.7860 4616 1.17 998.4 1.98 0.211 96.5
IP3J70 0.6877 4020 1.45 998.5 1.97 0.185 96.1
IP3J60 0.5895 3937 1.90 997.1 2.12 0.148 94.9
IP3J50 0.4912 3251 2.04 997.2 2.11 0.124 90.2
IP3J40 0.3930 2625 2.63 997.1 2.12 0.098 97.3
IP3J30 0.2947 1710 3.75 998.5 1.96 0.080 96.5
2D deposition profile measurement
IP3J80-2D 0.7860 5062 1.21 997.5 2.12 0.197 95.7
IP3J60-2D 0.5895 3893 1.74 997.2 2.16 0.145 97.5
IP3J40-2D 0.3930 2560 2.06 997.4 2.13 0.098 98.8

See Table 4.3 for parameter definition.

79
Table 4.5: Horizontal momentum jet experiments using MF particles, measuring
1D and 2D bottom depositions.
Run u0 Re C0 a ws lm RM
No. (m/s) (g/L) (kg/m3 ) (cm/s) (m) (%)
1D deposition profile measurement
MFJ70 0.6877 4981 0.484 996.09 2.33 0.157 89.8
MFJ60 0.5895 3946 0.687 997.05 2.31 0.136 88.1
MFJ50 0.4912 3273 0.825 997.13 2.22 0.118 88.8
2D deposition profile measurement
MFJ80-2D 0.7860 5097 0.473 997.43 2.18 0.191 92.4
MFJ60-2D 0.5895 4197 0.621 996.32 2.22 0.141 89.4
MFJ40-2D 0.3930 2572 0.806 997.33 2.20 0.095 98.0

See Table 4.3 for parameter definition.

4.6 Cross-sectional sediment concentration mea-


surement

Cross-sectional sediment concentration is measured for plastic (IP3) particle-


laden horizontal momentum jet using the particle imaging technique developed
by Lee (2010) and Lee et al. (2012). The principle of the method is to count the
number of particles (as bright spots) on high speed images based on the candi-
date particles size and light intensity. A 2mm thick laser sheet was generated by
an Argon-ion laser to cut across specific locations of the jet, using the same appa-
ratus described in Fig. 4.1. Scattered light from particles passing the laser sheet
was captured by a high-speed CMOS camera (PCO 1200 hs). Image to object
size ratio of 0.04 and f -number (ratio of focal length to aperture size) f # = 2 are
used for the experiments. Camera exposure time is 2 ms, less than the minimum
travel time of a particle across the laser sheet thickness. Each experimental run
for taking images for a cross-section is typically 3-4 minutes. Images capturing
begins 1 minute after start of an experiment. The image capturing interval is set
as 0.02 s and about 3400 images with 8-bit resolutions are recorded for about 70s.
The present experiments for particle concentration measurement is summarized
in Table 4.6.
Captured digital images are then analysed using a particle counting algorithm
developed in MATLAB programming language (Lee, 2010). The recognition of
candidate particles depends on the threshold light intensity value and the particle
image diameter. Particle image diameter can be estimated by optical theory (Lee,
2010). The particle image intensity value mainly depends on particle reflectance
properties, camera optical settings and laser intensity. Direct measurement such
as suction measurement can be used to calibrate the threshold intensity value.
Alternatively in a horizontal sediment-laden jet it can be calibrated by checking
mass balance with the bottom deposition rate data (Lee, 2010), i.e. the sum of
the sediment mass flux across a cross section Fx at xk and the total sediment

80
deposition flux Fb before xk should be equal to the jet sediment mass flux Fh
(Fig. 4.7), i.e.
Z xk Z
Fh = Fb (xk ) + Fx (xk ) = Fs (x)dx + u(x)c(x)dA (4.8)
0 A

where Fs is the sediment deposition flux (g/m/s) measured in bottom deposition


experiments.
In order to apply the same threshold intensity value at different jet flows and
cross sections, the camera optical settings are set as the same for all experiments
(Lee, 2010). A dimensionless threshold intensity parameter = Ith /Ip is intro-
duced where Ith is the threshold intensity value and Ip is the peak intensity value
( 242) after removal of the background noise. is found by trial and error to
best-match the mass balance with bottom deposition. The threshold intensity
parameter used for IP3 particles is 0.25. Particle overlapping is negligible for the
dilute particle concentrations in the present experiments.
The jet cross section with y, z coordinates [R, R; R, R] are divided into a
9 9 grid, where R = 2bT , bT = 0.16x = top-hat jet half-width. The number
of particles in each grid in each image is counted. The sediment concentration
Ci in each grid is determined by multiplying the average particle number by
the mass of each particle m = d3 /6 and dividing by the measurement volume
Vi = xyz, where x = 2mm is the laser sheet thickness, y and z are
the grid sizes = 0.4bT .

Table 4.6: Cross-sectional sediment concentration measurement, IP3 particles.


Run u0 Re C0 a ws lm Location of transverse
No. (m/s) (g/L) (kg/m3 ) (cm/s) (m) profiles (x/D)
IP3J80-C 0.7860 4616 1.17 998.4 1.98 0.211 12.5, 20.8, 33.3, 50.0, 66.7
IP3J50-C 0.4912 3251 2.04 998.4 1.98 0.124 8.3, 12.5, 20.8, 33.3, 50.0

See Table 4.3 for parameter definition.

81
Figure 4.7: Calibration of sediment concentration measurement by mass balance
between the jet cross section and bottom deposition.

4.7 Measurement of jet turbulent velocity

4.7.1 Particle Imaging Velocimetry

Experiments are carried out to study the loitering effect assumption in Nielsens
Lagrangian auto-correlation function. The horizontal and vertical velocities of a
turbulent jet are measured for four pure single phase jet cases (Qj = 40, 50, 60, 80
L/h) using the same experimental set up using high measurement frequency (100-
200Hz). Particle Imaging Velocimetry (PIV) is used for the velocity measure-
ment. PIV is a whole-field velocity measurement technique which utilizes the
correlation between successive images to determine the Eulerian velocity of each
interrogation window in the image (Raffel et al. 1998). It is a non-intrusive and
much more efficient technique for determining a large number of 2D velocity vec-
tors at the same time, compared to the tedious point measurement using Laser
Doppler Anemometry (LDA) or the intrusive point measurement of Acoustic
Doppler Velocimetry (ADV).
The experimental set up is the same as the sediment-laden jet, only with-
out sediment added. Before the experiment, neutrally buoyant (d = 50m ,
p = 1.003g/cm3 ) seeding particles are added to the source and ambient fluid.
A laser sheet is used to illuminate seeding particles at the jet centerline plane
in the vertical direction. The laser sheet is generated using an Argon ion con-
tinuous laser with a series of mirror and a cylindrical lens. Although the power
(light intensity) of continuous Argon-ion laser is considerably lower than a pulsed
Nd:YAG laser typically used in many PIV studies, the pulse laser available is not
able to generate pulses in high frequency of 100-200 Hz for the continuous turbu-
lence measurement. Such high frequency measurement has to be achieved by the
delay of a high-speed camera. Images of the experiments are captured by a 10

82
bit high speed CMOS camera (PCO 1200 hs, maximum frequency 500Hz) with 2
Gigabyte memory, resolution of 1280x1024 pixel. Before the experiment, a grid
is placed in the water tank to adjust the camera focus and calibrate the cameras
pixel to actual size ratio for later analysis. The exposure time of each image is set
as 2-3 ms depending on the flow velocity to freeze the particles in the image.
The time interval for each photo is 5ms for Qj = 80L/hr (u0 = 0.79m/s), 7.5ms
for Qj = 50 and 60L/hr (u0 = 0.49 and 0.59 m/s) and 10ms for Qj = 40L/hr
(u0 = 0.39m/s). A total of over 2300 images can be captured continuously. De-
pending on the jet flow rate, a total duration of 25-50s of turbulent flow data
can be captured. For each image, the coverage is 0-60D (about 0.4m) in the
longitudinal direction and 20D (about 0.12m) in the vertical direction.
After the experiment, the images are analysed using the Dantec FlowManager
PIV software (Dantec Dynamics, 2000). The image is divided into a number of
interrogation windows of 32 (horizontal) by 16 (vertical) pixels an overlapping
ratio of 50%. Vectors are analysed from the image by the software. Raw vectors
after correlation analysis were subjected to a screening procedure - to filter out
unrealistic and excessively large or small vectors. The filtering was based on the
maximum velocity at the source, i.e. it is impossible to get velocities at any point
larger than the exit velocity u(x, z) u0 . Over 90% of the vectors are considered
as valid in the measurement. Any stray vector is replaced by the average of the
eight vectors surrounding it. Only the results starting from x = 20D (0.12m) is
used for analysis as the spatial resolution is insufficient to resolve the flow near
the jet nozzle.

4.7.2 Mean flow and turbulent fluctuations


The measured mean flow velocities are shown in Fig. 4.8 for centerline and Fig. 4.9
for cross-sectional profiles. They compare well with the classical analytical solu-
tion for a simple jet.
 
uc (x) x 1
= 6.2 (4.9)
u0 D
!
u(x, r) r2
= exp 2 (4.10)
uc (x) b
Figs. 4.10 and 4.11 show the RMS turbulent velocity fluctuations normalized
with the centerline axial velocity ( u u /uc , w w /uc ). The measured turbulent

intensity has a peak of about 0.25 at the centerline for axial velocity fluctuations
and about 0.2 for radial velocity fluctuations. The measurement compares well
with the past studies (e.g. Papanicolaou and List, 1988; Wang and Law, 2002)
and correspond well to the CFD prediction using realizable k turbulence model
(see Fig. 5.7, Chapter 5). The normalized turbulent profiles are similar for the
experiments with a wide range of jet initial velocity, indicating the applicability of
a single turbulence intensity profile in the numerical particle tracking model. The
experimental data also shows that the current set-up using continuous Argon-ion
laser is capable of measuring the turbulent velocity with high accuracy and the
velocity data can be further utilized for analyses of turbulence characteristics of

83
jet flow. The correlations between |u | and |u /t|, and |w | and |w /t| for
the experimental support of the loitering effect are presented in Chapter 3.
1 1
uc/u0 uc/u0
Measured
Analytical

0.1 0.1

J40 x/D J80


0.01 0.01
1 10 100 1 10 x/D 100

(a) Q j = 40L/h, u0 = 0.39m/s (b) Q j = 80L/h, u0 = 0.79m/s

Figure 4.8: Comparison of measured and analytical axial centerline velocity uc


(Eq. 4.9).

1.2 1.2
u/uc 20D u/uc 60D
30D 40D
1 40D 1 50D
50D 30D
60D 20D
0.8 Gaussian 0.8 Gaussian

J40
J80
0.6 0.6

0.4 0.4

0.2 0.2

0 0
-3 -2 -1 0 1 2 r/b 3 -3 -2 -1 0 1 2 r/b 3

(a) Q j = 40L/h, u0 = 0.39m/s (b) Q j = 80L/h, u0 = 0.79m/s

Figure 4.9: Comparison of measured and analytical cross-sectional axial center-


line velocity u(x, r)/uc (x), Eq. 4.10.

84
0.4 0.4
20D 20D
u ' u' 30D
u ' u'
0.35 0.35 30D
uc 40D uc 40D
50D 50D
0.3 0.3
60D 60D

0.25 J40 0.25 J80

0.2 0.2

0.15 0.15

0.1 0.1

0.05 0.05
r/b
0 0
-3 -2 -1 0 1 2 r/b 3 -3 -2 -1 0 1 2 3

(a) Q j = 40L/h, u0 = 0.39m/s (b) Q j = 80L/h, u0 = 0.79m/s



Figure 4.10: Axial RMS turbulent fluctuations u u /uc .

0.4 0.4

w' w' J40 w' w' J80


0.35 0.35
20D
uc 30D
uc
0.3 40D 0.3
50D
0.25 60D 0.25

0.2 0.2

0.15 0.15

0.1 0.1

0.05 0.05

0 0
-3 -2 -1 0 1 2 r/b 3 -3 -2 -1 0 1 2 r/b 3

(a) Q j = 40L/h, u0 = 0.39m/s (b) Q j = 80L/h, u0 = 0.79m/s



Figure 4.11: Radial (vertical) RMS turbulent fluctuations w w /uc .

85
4.7.3 Experimental support of trapping effect in jet tur-
bulence
Nielsens Lagrangian auto-correlation function for a sediment particle consists
of an important assumption that generate the so called loitering effect which
results in the reduction of settling velocity - the assumption of slow velocity
corresponds to slow change of velocity. This represents a turbulent structure
that when a particle reaches a region of slow velocity, the particle would be
trapped with reduced velocity as the change of velocity w is slow. With this
assumption the particle tracking model works well in predicting the deposition
of horizontal turbulent sediment laden jets. However, as remarked by Nielsen
(1992), the assumption has not been studied in any types of flow.
Such an assumption can be supported if it can be shown a positive correlation
exists between the velocity magnitude |u | and the absolute value in the change
in velocity |u /t|. The basis of this assumption is studied using high frequency
PIV measurements of turbulent flow velocities of a horizontal jet. The horizontal
and vertical velocities of a turbulent jet are measured in a longitudinal vertical
plane. Fig. 4.12 shows an example of the measured time history of turbulent
velocity fluctuations. The change in turbulent velocity is estimated using forward
difference, e.g. !
w (w wi )
= i+1 (4.11)
t i t
The correlations between |u | and |u /t|, and |w | and |w /t| are analysed.
The measured correlations of |u | and |u /t|, and |w | and |w /t| are
presented in Figs. 4.13 and 4.14 for two jet flow cases. There is a significant
correlation between |w | and |w /t|, from about 0.3 at x = 30D to 0.6 at
x = 50D along the jet centerline (Fig. 4.13). On the other hand the correlation
between |u | and |u /t| is much less significant. Also, along the shown cross-
sections, the correlation of |w | and |w /t| is significantly higher (Fig. 4.14).
This gives the first direct experimental evidence that the assumption of slow
velocity corresponds to slow change of velocity is valid in a jet. The trapping
and loitering effect in particularly important and governing the sediment fall-out
in horizontal jets.

86
0.1
u' (m/s)

0.05

-0.05

t (s)
-0.1
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

(a) axial turbulent velocity


0.1
w' (m/s)

0.05

-0.05

t (s)
-0.1
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

(b) vertical (radial) turbulent velocity


10
2
u'/ t (m/s )
5

-5

t (s)
-10
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

(c) axial turbulent velocity acceleration


10
w'/ t (m/s ) 2

-5

t (s)
-10
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

(d) vertical (radial) turbulent velocity acceleration

Figure 4.12: Measured time history of turbulent velocity fluctuations (u , w ) and


its local acceleration (u /t, w /t). u0 = 0.39 m/s (Qj = 40 L/h), x = 30D,
z = 0, u = 0.080 m/s, w = 0.0014 m/s. Measurement interval = 0.01s.

87
0.9 u' u' / t w' w' / t u0 = 0.49m/s
,
u' u' / t w' w' / t z = 0 (centerline)
0.7

0.5

0.3

0.1 x/D

-0.1 0 10 20 30 40 50 60 70
u'
-0.3
w'
-0.5
(a) Qj = 40 m3 /s, u0 = 0.49 m/s

0.9 u' u' / t w' w' / t u0 = 0.79m/s


,
u' u' / t w' w' / t z = 0 (centerline)
0.7

0.5

0.3

0.1 x/D

-0.1 0 10 20 30 40 50 60 70
u'
-0.3
w'
-0.5
(b) Qj = 80 m3 /s, u0 = 0.79 m/s

Figure 4.13: The correlation between |u | and |u /t|, |w | and |w /t|, at jet
centerline (z = 0).

88
u ' u ' / t w' w' / t u ' u ' / t w' w' / t
, ,
u ' u ' / t w' w' / t u ' u ' / t w' w' / t
x = 30D 0.9 x = 30D 0.9

0.7 0.7

0.5 0.5

0.3 0.3

0.1 z/bg 0.1 z/bg

-3 -2 -1 -0.1 0 1 2 3 -3 -2 -1 -0.1 0 1 2 3
u' u'
-0.3 w' -0.3 w'

-0.5 -0.5
r/b r/b

u ' u ' / t w' w' / t u ' u ' / t w' w' / t


, ,
u ' u ' / t w' w' / t u ' u ' / t w' w' / t

x = 50D 0.9 x = 50D 0.9

0.7 0.7

0.5 0.5

0.3 0.3

0.1 z/bg 0.1 z/bg

-3 -2 -1 -0.1 0 1 2 3 -3 -2 -1 -0.1 0 1 2 3
u' u'
-0.3 w' -0.3 w'

-0.5 -0.5
r/b r/b
3
(a) Qj = 40 m /s, u0 = 0.49 m/s (b) Qj = 80 m3 /s, u0 = 0.79 m/s

Figure 4.14: The correlation between |u | and |u /t|, |w | and |w /t|, along
vertical transects at x = 30D and x = 50D

89
4.7.4 Experimental support of negligible cross-correlations
between partial velocity derivatives
The second assumption of the autocorrelation function: the cross-correlations

between the velocity components (u , v , w ) and their partial derivatives u
t
,
u u u
u x , v y , w z are zero, can be substantiated by the same data of jet velocity
measurement by PIV. Expanding the square of the partial derivative of velocity
components, it can be written as:
u component:
!2
u u u u
+u +v +w (4.12)
t x y z
!2 !2 !2 !2
u 2 u 2 u 2 u
= +u +v +w
t x y z
| {z } | {z }
Type 1 Type 2
! ! ! ! ! !
u
u
u
u
u
u
+2 u +2 v +2 w
t x t y t z
| {z }
Type 3
! ! ! ! ! !
u u u u u u

+2 u v +2 u w +2 v w
x y x z y z
| {z }
Type 4

v component:
!2
v v

v

v

+u +v +w (4.13)
t x y z
!2 !2 !2 !2
v v 2 2 v

2 v

= +u +v +w
t x y z
! ! ! ! ! !
v
v

v
v

v v

+2 u +2 v +2 w
t x t y t z
! ! ! ! ! !
v v v v v v

+2 u v +2 u w +2 v w
x y x z y z

w component:
!2
w w w w
+ u + v + w (4.14)
t x y z
!2 !2 !2 !2
w w w w
= + u2 + v 2 + w2
t x y z
! ! ! ! ! !
w
w

w
w

w w

+2 u +2 v +2 w
t x t y t z

90
! ! ! ! ! !
w w w w w w

+2 u v +2 u w +2 v w
x y x z y z

There are four combination of partial derivatives after the squaring:



1. The square of time derivatives: e.g. ( u
t
)2 , ( w
t
)2 ,

2. The square of spatial derivatives: e.g. (u u
x
)2 , (w u
z
)2 ,

3. The product of time and spatial derivatives: e.g. ( u
t
)(u u
x
), ( u
t
)(w u
z
),
and

4. The product of spatial derivatives: e.g. (u u
x
)(w u
z
).

Types 1 and 2 are square terms thus their mean are non-zero. Types 3 and 4
arises from the cross multiplication due to the squaring of the partial derivative
thus can be positive or negative. By examing the velocity measurement using
PIV, the magnitude of the mean value of these four types of partial derivatives
can be compared for their relative importance. Because of the axisymmetric
nature of a round jet, the v and w are taken as the same statistically. In
estimating the partial derivatives, the spatial derivatives are computed using a
central difference scheme centered at the interrogation window, while the time
derivative is computed using a forward difference scheme.
Table 4.7 shows the magnitude of the four types of velocity derivatives for
the jet case of Q = 50 L/h (u0 = 0.492 m/s) at the jet centerline z = 0. The
major contribution to the square of the partial derivative is Type 1: the time
derivative, with an order of magnitude of 1 m2 /s4 . Types 2 and 3 have a much
less significant contribution of the order of 104 103 m2 /s4 . Type 4 is negligible
with the order of 107 106 m2 /s4 . Table 4.8 shows the cross-correlations of
the velocity derivatives for the jet case of Q = 50 L/h (u0 = 0.492 m/s) which
are all less than 0.1, substantiating the assumption that the cross-correlations
can be neglected.

Table 4.7: Time average of the velocity derivatives for the jet case of Q = 50
L/h (u0 = 0.492 m/s). D = jet diameter = 6mm.
Velocity derivatives Type Time average (m2 /s4 , N = 2325)
x = 30D x = 40D x = 50D x = 60D
(u /t)2 1 2.30E+00 1.58E+00 9.71E-01 9.53E-01
(u u /x)2 2 1.08E-03 3.25E-04 1.54E-04 1.28E-04
(w u /z)2 2 1.67E-03 5.75E-04 1.91E-04 4.13E-05
(u /t)(u u /x) 3 2.85E-03 -1.42E-03 3.09E-04 4.32E-04
(u /t)(w u /z) 3 -2.06E-03 1.02E-03 2.71E-04 -2.65E-04
(u u /x)(w u /z) 4 3.77E-05 -6.40E-07 -3.50E-06 2.45E-07

91
Table 4.8: Velocity derivatives and the result of correlation for the jet case of
Q = 50 L/h (u0 = 0.492 m/s). D = jet diameter = 6mm.
Velocity derivatives Correlation results (z = 0)
Time
D  
derivatives
E D E D E
x = 30D x = 40D x = 50D x = 60D
u u u u
u / u 0.057 -0.063 0.025 0.039
D t   x E D t E D x E
u
t
w uz
/ u
t
w uz
0.035 0.034 0.020 0.010
D  E D ED E
w
u w / w u w -0.061 -0.066 0.023 -0.017
D t   x E D t E D x E
w
t
w w
z
/ w
t
w w
z
0.039 0.016 0.047 -0.060
Spatial
D
derivatives

(x zEplane)
E D D E
30D 40D 50D 60D
u u
u
u x w z / u u x
w z
0.026 -0.002 -0.022 0.003
D

E D
ED
E
u w
x
w w
z
/ u w
x
w w
z
0.012 -0.034 0.038 0.044

4.8 Summary
In this chapter, the experiments of sediment-laden horizontal jets are described.
Twenty-one sediment-laden jet experiments are carried out to measure the 1D
and 2D bottom deposition profiles using three types of particles of different size
and density, with flow visualization recorded. Checking of the mass balance
between system input and output showed excellent mass recovery percentage.
Cross-sectional sediment concentration of plastic particle-laden jets (two cases)
is measured using particle imaging technique. The experimental data will be
used for validating the numerical models and presented in Chapter 5.
Measurement of single phase jet velocity using PIV technique is carried out
for experimental support of the autocorrelation function developed in Chapter 3.
The measured mean flow and RMS turbulent velocity show very good comparison
with the analytical solution and results of past investigations. The positive
correlation between the radial (vertical) velocity magnitude and the absolute
change in velocity supports the assumption in the autocorrelation function for
modelling the loitering and trapping of sediment particles in jet turbulent eddies.

92
Chapter 5

Modelling of Sediment-Laden
Horizontal Momentum Jets

5.1 Introduction
The particle tracking model developed in Chapter 3 aims to predict the sediment
transport and deposition of horizontal momentum jet. In this chapter, the results
of experimental and numerical modelling of sediment-laden horizontal momen-
tum jet are presented. First, the experimental results of the present and previous
studies are discussed. As a preliminary investigation tool, Computational Fluid
Dynamics (CFD) modelling is utilized to study the interaction between parti-
cle and turbulence in horizontal momentum jets. Finally, the results of particle
tracking model prediction is compared with the experimental data of sediment
deposition and concentration. The importance of various forces in the equation
of motion is discussed through comparison with experimental measurements.

5.2 Horizontal sediment-laden momentum jet


Fig. 5.1 shows a schematic depiction of the structure of a horizontal sediment-
laden momentum jet with dilute concentration. The jet with a diameter D and
initial velocity u0 mixes with ambient fluid by shear-induced turbulent entrain-
ment. Sediment particles with source concentration C0 are transported in the
horizontal direction and dispersed by turbulent mixing. As the flow velocity and
turbulence intensity reduce, particles fall to the bottom under gravity, forming
a deposition profile with a peak close to the jet nozzle and an elongated tail.
For the present experiments, the sediment concentration is less than 7.7 g/L,
corresponding to a volume fraction of less than 0.29%. Cuthbertson et al. (2008)
showed experimentally that for sediment volumetric fraction of 0.1%, the jet
trajectory is unaffected by the presence of the sediment. Consistent with the
present experiments, the dilute particle concentration is considered to have little
influence on the fluid jet flow.

93
Turbulent jet

u0, C0

Particle
Fallout
Sediment
Deposition

Figure 5.1: Experiment of horizontal sediment-laden jets

5.3 Experimental observation

5.3.1 Flow visualization


To qualitatively study the dynamics of horizontal sediment-laden jets, high-speed
images are taken for a sediment-laden jet on a vertical laser sheet plane along
the jet flow direction. Fluorescent Rhodamine-6G dye is added to the jet flow
for the visualization of the fluid phase. As the exposure time of each photo is
very short (1-2 ms depending on jet flow), the particles are essentially frozen
on the image.
Fig. 5.2a shows a reversed gray-scale image of the experimental case of G180J80
(u0 = 0.786 m/s, glass particle d50 = 180m, ws = 2cm/s, C0 2.4g/L). The
turbulent nature of the jet fluid flow can be clearly observed with the turbulent
eddies visible under the fluorescent dye. Particles at first travelled along the
direction of the jet flow. The particles are seen to be bursting upward and down-
ward with the turbulent eddies. At some distance, particles starts to drop out
from the jet flow as the mean flow velocity reduces and turbulent intensity weak-
ens. The number of sediment particles at the upper part of jet starts decreasing
and particles are concentrated at the lower half of the jet. Further downstream,
the sediment in the turbulent jet has been reduced to low numbers that almost
only the turbulent fluid phase remains. Fig. 5.2b shows the visualization of a
pure jet of the same initial velocity. The jet centerline and top-hat width, defined
by bT = 0.16x, is included for illustration. Compared to the pure jet case, the

94
fluid phase of the sediment-laden jet is similar to the pure jet one, indicating that
the dynamics of the jet flow is not significantly affected by the presence of sed-
iment. The sediment volume fraction for this case is 0.096%; the visualization
confirms the conclusion by Cuthbertson et al. (2008) that when the sediment
volume fraction is less than 0.1%, the jet flow is not significantly affected by the
presence of sediment.
Fig. 5.3a shows the visualization of the experimental case of G180J50 (u0 =
0.492 m/s, glass particle d50 = 180m, ws = 2cm/s, C0 3.8g/L), which the
jet initial velocity is smaller than the previous case. Because of the lower jet
velocity, sediment falls out closer to the nozzle, leaving a clean fluid phase behind
downstream. Higher sediment concentration is used in this case (volume fraction
0.152%) which is slightly higher than the limit of no-impact by Cuthbertson et
al. (2008). Nevertheless, the mean flow and turbulent structures of the sediment-
laden jet is similar to those of the pure jet (Fig. 5.3b).
Fig.5.4 shows the visualization of a jet with u0 = 0.786m/s and laden with
plastic IP3 particles of 716 m diameter (ws = 2.1cm/s, p = 1140 kg/m3 ).
For the plastic particle experiments, the initial concentration of sediment are
lower in order to cater for the larger particle size. Due to the low concentration
(vol. fraction 0.1%), the trajectory of the jet fluid phase is unaffected by the
presence of the particle (Cuthbertson et al. 2008). The particles travel with the
turbulent eddies and fall out from the lower jet edge but with a longer distance
compared to the glass particles.

5.3.2 Sediment bottom deposition


One- and two-dimensional sediment deposition profiles are measured in the present
study using various particles. The experiments are summarised in Chapter 4. In
addition, extensive 1D sediment deposition has been reported in Li (2006) and
Lee (2010) for a large number of horizontal momentum sediment-laden jet exper-
iments using natural sand and glass particles respectively, and limited number of
experiments using plastic particles. A summary of the experimental parameters
for their work is given in Table 5.1.

1D (transversely summed) deposition profiles

The 1D longitudinal deposition rate is obtained by summing the deposited mass


in the transverse y-direction. Fig. 5.5 shows typical measured 1D bottom depo-
sition profiles from horizontal momentum jets with glass and plastic particles.
The deposition is initially close to zero near the jet nozzle as sediment has not
begun falling out from the jet edge. At some distance particle deposition appears
and rises to a maximum at a distance of xm which correspond to the maximum
rate of falling out from the jet. Then the deposition drops with the shape of a
longer tail to zero, as the sediment concentration in the jet decreases due to fall
out. The higher the jet flow velocity, the longer the deposition pattern is, as the
sediment particles are transported further downstream.

95
0.08

0.04
(m)

0.04

0.08
0 0.08 0.16 0.24 0.32
(m)
(a) Sediment laden jet, glass particle d50 = 180m , ws = 2cm/s

0.08

0.04
(m)

0.04

0.08
0 0.08 0.16 0.24 0.32
(m)
(b) Pure jet

Figure 5.2: Visualization of case u0 = 0.786 m/s, (a) sediment laden, (b) pure
jet. The dashed line is the jet centerline; the dash-dotted line is the top-hat
width of the jet, defined by bT = 0.16x.

96
0.08

0.04
(m)

0.04

0.08
0 0.08 0.16 0.24 0.32
(m)
(a) Sediment laden jet, glass particle d50 = 180m , ws = 2cm/s

0.08

0.04
(m)

0.04

0.08
0 0.08 0.16 0.24 0.32
(m)
(b) Pure jet

Figure 5.3: Visualization of case u0 = 0.492 m/s, (a) sediment laden, (b) pure
jet. The dashed line is the jet centerline; the dash-dotted line is the top-hat
width of the jet, defined by bT = 0.16x.

97
0.08

0.04

0
(m)

0.04

0.08

0 0.08 0.16 0.24 0.32


(m)

Figure 5.4: Visualization of IP3J80, u0 = 0.786 m/s, plastic IP3 particle d50 =
716m , ws = 2.2cm/s

Fs (g/m/s) Fs (g/m/s)
0.5 0.1

G180J50 0.09 G115J40


0.4 G180J80 0.08 G115J60
0.07
0.3 0.06
0.05
0.2 0.04
0.03
0.1 0.02
0.01
x (m)
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1

(a) Glass, d50 = 180m, ws = 2.0 cm/s (b) Glass, d50 = 115m, ws = 1.0 cm/s
(Lee 2010) (Lee 2010)
Fs (g/m/s) Fs (g/m/s)
0.2 0.06
IP3J80 MFJ80
0.18
IP3J60 0.05 MFJ60
0.16
IP3J40 MFJ40
0.14
0.04
0.12
0.1 0.03
0.08
0.02
0.06
0.04
0.01
0.02
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1

(c) IP3, d50 = 716m, ws = 2.2 cm/s (d) MF, d50 = 347m, ws = 2.2 cm/s
(present study) (present study)

Figure 5.5: Typical 1D sediment deposition profiles (present study and Lee,
2010).

98
Table 5.1: Summary of Li (2006) and Lee (2010)s horizontal sediment-laden jet
experiments for bottom deposition and cross-sectional sediment concentration
measurement. Jet diameter D = 6mm. lm = momentum-settling length scale
1/2
= M0 /ws .
Particle Particle Settling Jet Jet Sediment Momentum/
Type diameter velocity flow rate velocity concentration settling
daeq /db50 ws Q0 u0 C0 length scale
(m) (cm/s) (L/hr) (m/s) (g/L) lm /D
3
Li (2006), sediment bottom deposition measurement, p = 2.65g/cm
Coarse 166a 1.98 50,54,58, 0.49 - 0.65 3.39 - 4.49 22.2 - 29.2
sand 62,66
(CJ)
Fine 133a 1.41 30,34,38, 0.30 - 0.57 3.93 - 7.73 18.5 - 35.7
sand 42,46,50,
(FJ) 54,58
Lee (2010), sediment bottom deposition
and concentration measurement, p = 2.5g/cm3
Spherical 215b 2.64 60,70, 0.57 - 0.86 3.39 - 4.49 18.9 - 28.3
Glass 215m 80,90
(G215J)
Spherical 180b 1.83 50,60, 0.48 - 0.76 2.32 - 3.71 20.6 - 32.9
Glass 180m 70,80
(G180J)
Spherical 115b 1.03 40,50, 0.38 - 0.67 1.03 - 1.83 32.9 - 57.6
Glass 115m 60,70
(G115J)
Lee (2010), sediment bottom deposition measurement, p = 1.14g/cm3
Spherical 716b 2.06 50,58, 0.48 - 0.63 1.56 - 2.26 18.5 - 24.8
Plastic 66
(IP3J)
Lee (2010), sediment bottom deposition measurement, p = 1.16g/cm3
Irregular 634a 2.11 50,58, 0.48 - 0.63 1.03 - 1.83 20.2 - 26.1
Plastic 66
(PJ)

99
For a horizontal sediment-laden jet, the sediment deposition pattern and
concentration profiles depends on the relative importance of the jet momentum
and sediment settling velocity. Based on the heuristic assumption that particles
start to settle when the settling velocity exceeds the jet entrainment velocity
ve = uc , where is the entrainment coefficient and uc is the centerline mean
flow velocity (Bleninger et al. 2002; Li, 2006 and Lee, 2010), a settling-momentum
length scale lm can be defined,
1/2
M
lm = 0 (5.1)
ws
where lm is a measure of the distance at which sediment starts to fall out of the jet.
The range of lm /D is 19-58 in the experiments of present study. When x < lm ,
the horizontal transport of sediment by jet flow dominates over the settling while
the other way occurs when x > lm . By dimensional analysis, Li (2006) and Lee
(2010) have found that if the longitudinal coordinate x is normalized as x/lm ,
and the deposition rate is normalized by the jet sediment flux Fs lm /Q0 C0 , the
deposition profiles of experiments using similar types of particles collapse into a
single profile and can be described by a log-normal distribution, as
"  2 #
Fs lm x
= A exp B ln C (5.2)
Q0 C0 lm

where the constants A, B and C are to be determined by best-fitting.


Fig. 5.6 shows the normalized profiles of 1D deposition using natural sand and
glass particles, from the experiments of Li (2006) and Lee (2010). The stillwater
settling velocity ws is used to calculate the momentum-settling length scale lm .
The profiles collapse well into a single curve fitted by a log-normal function.
Figs. 5.7 and 5.8 show the normalized deposition profiles of sediment-laden
jets with plastic IP3 and MF particles measured in this study. It can be observed
that the profiles collapse onto a single curve for each kind of particles. Never-
theless, comparing the three profiles (Fig. 5.9), the curves for glass and plastic
particles cannot be collapsed into a single one despite their similar shapes. The
jet turbulence-particle interaction may be quite different for glass and plastic
particles, even for similar jet flow conditions. The stillwater settling velocity ws
is not the only governing factor for the deposition in horizontal sediment-laden
jets. A general model incorporating the physics of particle-turbulence interaction
is required for the correct prediction of the bottom deposition profiles.

2D Deposition Profiles

Fig. 5.10 shows the measured 2D deposition profiles of experiments using G180
particles. The profiles are consistent with the previous measured 1D profiles,
with a peak deposition at about 0.2m for G180J80-2D and 0.1m for G180J50-2D.
It is noted that the deposition profiles are not symmetric about the centerline
of the jet. This is probably due to the unstable entrainment flow by the jet
in a finite sized tank together with the falling out of the sediment particles.

100
1
Glass (Lee, 2010)

0.8 Sand (Li, 2006)

Fitted eqn. (Li, 2006)


0.6
Fs lm /Q0 C0

0.4

0.2

0
0 1 2 3 4 5
x/lm

Figure 5.6: Normalized 1D deposition profiles of glass (Lee, 2010) and natural
sand particles (Li, 2006)

1
IP3-Present Study
0.8
IP3-Lee (2010)

NQP-Lee (2010)
0.6
Fs lm /Q0C0

Fitted eqn (present)


0.4

0.2

0
0 1 2 3 4 5
x/lm

Figure 5.7: Normalized 1D deposition profiles of plastic IP3 particles (present


study)

101
1
Measured
0.8 Fitted eqn.

0.6
Fs lm /Q0C0

0.4

0.2

0
0 1 2 3 4 5
x/lm

Figure 5.8: Normalized 1D deposition profiles of melamine MF particles (present


study)

1
IP3 (present)

0.8 MF (present)

Sand (Li, 2006)


0.6
Fs lm /Q0 C0

0.4

0.2

0
0 1 2 3 4 5
x/lm

Constants for Eq. 5.2


Constants Glass and Sand IP3 MF
Li (2006), Lee (2010) present study present study
A 0.702 0.594 0.560
B 1.5 2.01 1.87
C 1.19 0.875 0.921

Figure 5.9: Normalized 1D deposition profiles (fitted equations, Eq. 5.2) of dif-
ferent particles

102
Such phenomenon can be observed in the cross-sectional sediment concentration
measurement in Lee (2010). For plastic particles, the deposition profile shows
similar features (Fig. 5.11).

0.06
G180J80
Observed
0.04
1 1
0.02 3 2
y (m)

0
6 4

5
1

1
-0.02 3 4 3
2
2

1
-0.04

-0.06
0 0.1 0.2 0.3 0.4 0.5 0.6
x (m)

(a) u0 = 0.786 m/s

0.06
G180J50
Observed
0.04
2
0.02
6
y (m)

0
2
6

-0.02 2

-0.04

-0.06
0 0.1 0.2 0.3 0.4 0.5
x (m)

(b) u0 = 0.492 m/s

Figure 5.10: 2D deposition profiles of G180 particles (present study). Contours


in g/m2 /s.

103
0.06
IP3J80
Observed
0.04
0.5 0.25
0.02
0.5 0.75

y (m)

5
0.2
0

1.25
0.75

5
1.

1
1
75
-0.02 0. 25

0.5
0.5 0.
-0.04 0.25

-0.06
0 0.1 0.2 0.3 0.4 0.5 0.6
x (m)

(a) u0 = 0.786 m/s

0.06
IP3J40
Observed
0.04

0.02
0.5 1.5 0.
5
y (m)

23 3 1
0
3.5

2
2.5
-0.02 1 1.5
0.5
-0.04

-0.06
0 0.1 0.2 0.3 0.4 0.5
x (m)

(b) u0 = 0.393 m/s

Figure 5.11: 2D deposition profiles of IP3 particles (present study). Contours in


g/m2 /s.

5.4 3D Computational Fluid Dynamics mod-


elling
3D Computation Fluid Dynamics (CFD) modelling is initially adopted for study-
ing the dynamics of sediment laden jets. The 3D Reynolds-averaged Navier-
Stokes (RANS) equations are solved numerically for the steady jet flows. The
following sediment transport equation commonly adopted in environmental and
engineering sediment transport modelling is solved to estimate the sediment con-
centration C. !
C t C
(Ui ws ) = (5.3)
xi xi Sct xi
where C is the sediment concentration and Sct is the turbulent Schmidt number
taken as 0.85. The term ws C z
is added for the settling flux of sediment, where
ws is the settling velocity. The CFD model is described in detail in Appendix B.
Fig. 5.12 shows the comparison of the sediment deposition profiles by labo-
ratory measurement and CFD for a number of experiment cases. When the still
water settling velocity is used in the Eulerian CFD model, the predicted deposi-
tion profile does not compare well with the experimental measurement. A higher
maximum deposition rate is predicted with its position closer to the jet nozzle.
This indicates the importance of jet turbulence in reducing the settling velocity.
If the settling velocity is reduced by 25% in the whole flow field, a much closer
comparison is obtained. This 25% reduction applies to all the experimental cases

104
using natural sand and glass particles (Fig. 5.13).
Another approach to adjust the settling velocity is to reduce it locally accord-
ing to the ratio of RMS turbulent fluctuation to terminal settling velocity /ws ,
based on the model predicted apparent settling velocity in grid turbulence in
Chapter 3, i.e. Eq. 3.44. Fig. 5.14a shows the /ws ratio in various cross-section
of a jet experiment case FJ42. Within two Gaussian width b the RMS turbulent
fluctuation is very high and nearly 4 times of the settling velocity. Fig. 5.14b
shows the apparent settling velocity estimated from Eq. 3.44. Near the centerline
of the jet, a reduction close to 30% of the terminal settling velocity is possible.
This value is consistent with the empirical reduction factor used by Li (2006)
for her jet integral model which is merely obtained by fitting with experimental
data. The reduction decreases as the radial position increases. When r > 2b, the
terminal settling velocity resumes as the jet turbulence no longer influences the
particle settling. The CFD predicted bottom deposition using this local adjust-
ment approach is shown in Figs. 5.12 and 5.13 and shows similar performance
compared to the whole field adjustment approach.
For the lower density plastic particles (Fig. 5.15), a further 10% reduction
is required to repredict the deposition profile correctly, although they have sim-
ilar stillwater settling velocity as the glass particles. This is consistent with
the prediction in homogeneous turbulence (Fig. 3.15) for which plastic particles
tends to settle much more slowly in high turbulent intensity. Thus the empirical
correction of settling velocity does not only depends on the stillwater settling
velocity and the turbulence level, but is also related to the particle properties
(size, density). The local adjustmust according to /ws ratio does not work well
for plastic particles (Fig. 5.15).
In summary, model-data comparison shows that a simple incorporation of
the settling flux using the still water settling velocity in CFD models is found
not to work well in the modelling of bottom deposition of horizontal sediment-
laden jets. Empirical adjustment of stillwater settling velocity is required for the
satisfactory prediction of the sediment deposition. A whole field adjustment
factor of 0.75 can be multiplied to the stillwater settling velocity to achieve
a better comparison with measured data for sand and glass particles, but for
plastic particles a smaller factor of 0.65 has to be used. A local adjustment factor
according to the /ws ratio can also be applied to give satisfactory prediction on
the deposition of sand and glass particles. The non-universality of the empirical
correction factor necessitates the development of a general model that does not
require any empirical a priori adjustment to the settling velocity for sediment-
laden jets, which will be described in next section.

105
0.5 0.5
CJ66 CJ50
0.4 CFD 0.75Ws CFD 0.75Ws
0.4
CFD Wsa CFD Wsa

Fs (g/m/s)
CFD Ws CFD Ws
Fs (g/m/s)

0.3 0.3
Meas. Meas.
0.2 0.2

0.1 0.1
x (m) x (m)
0 0
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8

(a) CJ66, u0 = 0.648 m/s, ws = 1.96 cm/s (b) CJ50, u0 = 0.491 m/s, ws = 1.96 cm/s
0.5 0.5
FJ58 FJ30
0.4 CFD 0.75Ws 0.4 CFD 0.75Ws
CFD Wsa CFD Wsa
Fs (g/m/s)
Fs (g/m/s)

0.3 CFD Ws 0.3 CFD Ws


Measured Measured
0.2 0.2

0.1 0.1
x (m) x (m)
0 0
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8

(c) FJ58, u0 = 0.570 m/s, ws = 1.41 cm/s (d) FJ30, u0 = 0.295 m/s, ws = 1.41 cm/s

Figure 5.12: CFD predicted and observed (Li, 2006) longitudinal particle depo-
sition pattern of four experiments using sand particles. The solid lines represent
the prediction using the adjusted settling velocity: 0.75ws whole field adjust-
ment; wsa local adjustment according to /ws . The dashed line is prediction
using the stillwater settling velocity ws .

106
0.7 0.7
G215J90 G215J60
0.6 0.6
CFD 0.75Ws CFD 0.75Ws
0.5 CFD Wsa 0.5 CFD Wsa
Fs (g/m/s)

Fs (g/m/s)
CFD Ws CFD Ws
0.4 0.4
Measured Measured
0.3 0.3

0.2 0.2

0.1 0.1
x (m) x (m)
0 0
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8

(a) G215J90, u0 = 0.884 m/s, ws = 2.66 cm/s (b) G215J60, u0 = 0.592 m/s, ws = 2.74 cm/s
0.8 0.8
G180J80 G180J50
CFD 0.75Ws CFD 0.75Ws
0.6 CFD Wsa 0.6 CFD Wsa
Fs (g/m/s)

Fs (g/m/s)

CFD Ws CFD Ws
Measured Measured
0.4 0.4

0.2 0.2

x (m) x (m)
0 0
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8

(c) G180J80, u0 = 0.786 m/s, ws = 1.95 cm/s (d) G180J50, u0 = 0.491 m/s, ws = 1.98 cm/s

0.1 0.1
G115J70 G115J40

0.08 CFD-0.75Ws 0.08 CFD 0.75Ws


CFD-Wsa CFD Wsa
Fs (g/m/s)

Fs (g/m/s)

0.06 CFD-Ws 0.06 CFD Ws


Measured Measured
0.04 0.04

0.02 0.02
x (m) x (m)
0 0
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8

(e) G115J70, u0 = 0.688 m/s, ws = 1.00 cm/s (f) G115J40, u0 = 0.393 m/s, ws = 1.04 cm/s

Figure 5.13: CFD predicted and observed (Lee, 2010) longitudinal particle de-
position pattern of six experiments using spherical glass particles. The solid
lines represent the prediction using the adjusted settling velocity: 0.75ws whole
field adjustment; wsa local adjustment according to /ws . The dashed line is
prediction using the stillwater settling velocity ws .

107
4.5
s/ws 1.2
wsa /ws
(a) (b)
4
1
3.5
3 x = 10D 0.8
x = 20D
2.5
x = 30D 0.6
2 x = 10D
x = 40D
x = 20D
1.5 x = 50D 0.4
x = 30D
1
0.2 x = 40D
0.5 x = 50D
0 0
0 1 2 r/b 3 0 1 2 r/b 3

Figure 5.14: (a) Ratio of RMS turbulent velocity to settling velocity (/ws ),
and (b) ratio of apparent settling velocity to terminal settling velocity (wsa /ws ),
in different cross-section of a jet, Jet case FJ42: deq = 133m, u0 = 0.41m/s,
ws = 1.41cm/s

0.25 0.25
IP3J66 IP3J50

0.2 CFD-0.65Ws 0.2 CFD-0.65Ws


CFD-0.75Ws CFD-0.75Ws
CFD-Wsa CFD-Wsa
Fs (g/m/s)
Fs (g/m/s)

0.15 CFD-Ws 0.15


CFD-Ws
Measured
Measured
0.1 0.1

0.05 0.05
x (m)
x (m)
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6

(a) IP3J66, u0 = 0.648 m/s, ws = 2.28 cm/s (b) IP3J50, u0 = 0.491 m/s, ws = 2.25 cm/s

Figure 5.15: CFD predicted and observed (Lee, 2010) longitudinal particle depo-
sition pattern of two experiments using plastic particles. 0.65ws , 0.75ws whole
field uniform adjustment of stillwater settling velocity ws ; wsa local adjustment
according to /ws .

108
5.5 Particle tracking modelling
Based on the velocity autocorrelation function developed in Chapter 3, a particle
tracking model is developed to predict the sediment concentration and bottom
deposition of horizontal sediment-laden jets. The model incorporates the ana-
lytical mean flow velocity (axial and radial) of a pure momentum jet, with the
best-fitted self-similar RMS turbulent fluctuations (turbulent kinetic energy) and
turbulent dissipation rate predicted using the realizable k turbulent closure
model in CFD solution for a jet.

5.5.1 Mean flow velocity


The theoretical Gaussian axial jet mean flow velocity is adopted in the model
and are given by
 1
uc (x) x
= 6.2 (5.4)
u0 D
!
u(x, r) r2
= exp 2 (5.5)
uc (x) b
where uc is the jet centerline velocity, u0 is the initial velocity, D is the initial jet
diameter and b = x is the Gaussian half width and = 0.114 is the jet linear
spreading rate. The mean transverse velocity of the jet is given by the following
expression deduced by applying continuity to the jet cross-section (Lee and Chu,
2003).
vr (r) (1 exp(r2 /b2 )) (/)(r2 /b2 ) exp(r2 /b2 )
= (5.6)
uc r/b
where ve = uc is the entrainment velocity at r = b; = 0.057 is the jet
entrainment coefficient. Fig. 5.16 shows the mean velocity profiles of a jet and
the comparison with CFD prediction.
Velocity gradient is required in the full model for the fluid accelerations
(duf /dt)f . Only the spatial derivatives based on the mean flow velocity (e.g.
u/x, u/y, etc) are modelled. It is found the the contribution of the force
due to fluid acceleration has negligible effect on the predicted deposition and
concentration profiles.

5.5.2 Turbulence quantities


q
Assuming isotropic turbulence, the RMS turbulent fluctuation = 23 k is ob-
tained from the turbulent kinetic energy k of a momentum jet using a k
turbulence closure model with CFD simulation of a pure jet. The length and
time scales TE and LE are obtained from Eqs. 3.21 and 3.20 (Chapter 3) using k
and the turbulent kinetic energy dissipation rate . The CFD model predicted
1/3
and are normalized with the mean jet properties as uc and (b)uc respectively,

109
and fitted with an equation with the form of
    
(r) r r
= C1 exp C2 ( C3 )2 + exp C2 ( + C3 )2 (5.7)
uc b b

and
    
((r)b)1/3 r r
= C4 exp C5 ( C6 )2 + exp C5 ( + C6 )2 (5.8)
uc b b

to provide their spatial variated functions. The empirical coefficients



C1 0.2006 C4 0.2458

C2 = 1.4147 , C5 = 1.2498
C3 0.6647 C6 0.6594

are obtained using a least-square best-fitting method. Fig. 5.17 shows the nor-
malized radial distribution of turbulence quantities in a jet. The RMS turbulent
velocity fluctuation profiles are corresponding very well to the experimental mea-
surement by Papanicolaou and List (1988) and Wang and Law (2002), with a
centerline turbulence intensity of about 0.22, within the range of the previous
studies. A slightly higher /uc and is obtained by the CFD model at x = 20D.
Inclusion of this effect in the particle tracking model is tested and found to be
insignificant for predicting the deposition pattern. Thus a single normalized
turbulent quantity profile is used for all longitudinal locations.

1.2 0.6
uc/u0 Qj = 42 L/h 10D vr/auc 10D
Qj = 42 L/h
1 20D 0.4 20D
30D 30D
0.8 0.2 40D
40D
50D 50D
0.6 0
Theoretical
Theoretical
0.4 -0.2

0.2 -0.4

0 -0.6
0 1 2 r/b 3 0 1 2 3 4 r/b 5

(a) (b)

Figure 5.16: Theoretical and CFD predicted jet mean velocity profiles, (a) Lon-
gitudinal velocity, Eq. 5.5, (b) transverse velocity (positive outwards), Eq. 5.6.

5.5.3 Particle size


In the full model, a single particle diameter d is used. For spherical particles,
due to their narrow size range, the diameter is taken as the median diameter d50
of the particles used in the experiments. For irregular particles, an equivalent

110
/uc 1/3
0.4 0.4 (b) /uc
10D 10D
20D 20D
0.3 30D 0.3 30D
40D 40D
50D 0.2 50D
0.2
Best-fit Best-fit

0.1 0.1
Qj = 42 L/h
Qj = 42 L/h
0 0
0 1 2 3 0 1 2 3
r/b r/b
(a) (b)

Figure 5.17: Turbulent velocity fluctuation and dissipation rate dissipation, pre-
dicted by CFD model and fitted with empirical equations, (a) RMS turbulent
velocity fluctuation, Eq. 5.7, (b) Turbulent energy dissipation rate, Eq. 5.8.

diameter deq is estimated based on the measured terminal settling velocity ws


and inferred from the settling velocity formula with spherical drag law,
3CD
deq = ws2 (5.9)
4g(s 1)
where CD = f (Re) = f (ws deq /) is the drag coefficient determined iteratively
using Eq. 3.3. This deq obtained is different from the one used in Li (2006) which
is obtained using the equation for estimating the stillwater settling velocity of
natural sand particles (Sousbly, 1997).
h i
ws = (10.362 + 1.049D3 )1/2 10.36 (5.10)
deq
" #1/3
g(s 1)
D = deq (5.11)
2
For both natural irregular sand and spherical particles, the drag law and added
mass of a spherical particle is used. Table 5.2 shows the inferred particle equiv-
alent diameter deq .

5.5.4 Computational procedures


The stochastic particle tracking is performed using Eqs. 3.1, 3.7 and 3.31 for
Full Model (Eqs. 3.32 and 3.34 for Simplified Model) in Chapter 3 with the
mean jet flow velocity determined from Eqs. 5.4-5.6, and turbulent quantities
from Eqs. 3.20, 3.21, 5.7 and 5.8. Np = 50, 000 particles are used for each
jet simulation to obtain the deposition profile. Each particle shares an equal
fraction of the total sediment mass used in the experiment. The total sediment
mass can be computed by the product Qj C0 Texp , Texp as the total duration of
the numerical experiment, which is set as 5 minutes. It should be noted that the

111
Table 5.2: The inferred particle equivalent diameter deq for non-spherical par-
ticles

Particle Measured settling Temperature Density deq


velocity ws (cm/s) ( C) ratio m
Coarse sanda 1.98 21 2.65 171
Fine sanda 1.41 23 2.65 135
Nan Qiang Plasticb 2.11 27 1.16 634
(NQP)
Melamine (MF)c 2.2 27 1.50 347
a Li (2006), b Lee (2010), c Present study

mass represented by each numerical particle in the simulations is not the same as
the actual mass of a sediment particle in the physical experiments. The particles
are initially distributed as a Gaussian profile and released at the end of the zone
of flow establishment (ZFE, x = 6.2D), and tracked until they reach the level
of the bottom tray (depends on the experiments: z = 15cm above the bottom
collection tray for present experiments and Lee (2010)s experiment, z = 7cm
for Li (2006)s experiment). The 1D longitudinal deposition rate (in g/m/s) are
obtained by summing all particles at the y-direction under intervals of 0.02m at
the x-direction and divided by the experiment time Texp . The 2D longitudinal
deposition rate (in g/m2 /s) are obtained by summing all particles within grid cell
size of x = 0.045m and y = 0.015m and divided by the experiment time Texp .
Particle tracking calculations have been performed for all experiments in Tables
4.3, 4.4, 4.5 (present study) and Table 5.1 (Li, 2006 and Lee, 2010) using a time-
step of 0.001s and selected results are shown in the next section. Sensitivity tests
have shown that even when a larger time-step up to 0.005s is used, the predicted
maximum deposition rates differ by less than 3% difference.

5.6 Particle tracking modelling results

5.6.1 1D sediment deposition profiles


Fig. 5.18 shows the comparison with particle tracking models for the natural
sand experiments by Li (2006). The particle tracking model predicts the 1D de-
position pattern very well without any a priori correction to the settling velocity
with turbulence. The autocorrelation function developed in Chapter 3 correctly
models the turbulent velocity fluctuations which accounts for the induced set-
tling velocity reduction. The model predicts similarly well for the deposition of
spherical glass particles (Fig. 5.19).
The predicted deposition profiles, using either full model or simplified model,
are similar for these cases using natural sand and glass particles with density
ratio about 2.5. By the assumption of the simplified model, the particle velocity

112
is the same as the fluid velocity except in the vertical direction. The similar per-
formance of both models indicates that these particles response almost instantly
to turbulent fluctuations. The simplified model provides a more efficient alterna-
tive for use in general engineering design purposes, without the need for solving
the full equation of motion. The simplified model will be extensively applied in
the deposition from sediment-laden buoyant jets in Chapter 6.

0.5 0.5
CJ66 CJ50
0.4 Measured 0.4 Full Model
Full Model Simplified model
Fs (g/m/s)

Fs (g/m/s)
0.3 Simplfied model 0.3 Measured

0.2 0.2

0.1 0.1
x (m)
0 0
0 0.2 0.4 0.6 x (m) 0.8 0 0.2 0.4 0.6 0.8

CJ66: u0 = 0.65 m/s, C0 = 3.39 g/L, CJ50: u0 = 0.49 m/s, C0 = 4.49 g/L,
deq = 171m, ws = 1.98 cm/s deq = 171m, ws = 1.97 cm/s

0.5 0.5
FJ58 FJ30
Full Model Full Model
0.4 0.4
Deposition rate (g/m/s)

Simplified Model
Simplified Model
Measured
Fs (g/m/s)

0.3 0.3 Measured

0.2 0.2

0.1 0.1

x (m)
0 0
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 x (m) 0.8

FJ58: u0 = 0.57 m/s, C0 = 3.93 g/L, FJ30: u0 = 0.29 m/s, C0 = 7.73 g/L,
deq = 135m, ws = 1.41 cm/s deq = 135m, ws = 1.41 cm/s

Figure 5.18: Comparison of measured and particle tracking predicted deposition


rate profiles of experiments using natural sand particles (Li, 2006).

The bottom deposition characteristic of jets with plastic particles differs sig-
nificantly from that of sand and glass particles. The full model predicts the de-
position profiles very well, while the simplified model predicts higher peak closer
to the nozzle (Figs 5.20 and 5.21). This is consistent with the CFD prediction
for which a further 10% reduction is required to match the prediction with the
observed deposition, and supported by the the greater reduction in settling veloc-
ity in homogeneous turbulence (Chapter 3). The unsatisfactory prediction of the
simplified model shows that due to particle inertia, the plastic particles are not
following the fluid completely. The prediction of melamine particles with density
ratio 1.5 also shows similar results (Fig. 5.22), with the full model predicts better
than the simplified model.

113
0.7 0.7
G215J90 G215J60
0.6 0.6
Full Model Full Model
0.5 Simplified Model 0.5 Simplified Model
Fs (g/m/s)

Fs (g/m/s)
0.4 Measured 0.4 Measured
0.3 0.3

0.2 0.2

0.1 0.1

0 0
0 0.2 0.4 0.6 x (m) 0.8 0 0.2 0.4 0.6 x (m) 0.8

G215J90: u0 = 0.88 m/s, C0 = 3.84 g/L, G215J60: u0 = 0.59 m/s, C0 = 5.51 g/L,
d50 = 215m, ws = 2.66 cm/s d50 = 215m, ws = 2.74 cm/s

0.7 0.7
G180J80 G180J60
0.6 Full Model 0.6
Full Model
0.5 Simplified Model 0.5 Simplified Model
Measured
Fs (g/m/s)

Fs (g/m/s)

Measured
0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1
x (m)
0 0
0 0.2 0.4 0.6 x (m) 0.8 0 0.2 0.4 0.6 0.8

G180J80: u0 = 0.79 m/s, C0 = 4.94 g/L, G180J60: u0 = 0.57 m/s, C0 = 6.18 g/L,
d50 = 180m, ws = 1.95 cm/s d50 = 180m, ws = 1.82 cm/s

0.2 0.2
G115J60 G115J40
Full Model Full Model
0.15 0.15
Simplified Model Simplified Model
Fs (g/m/s)

Fs (g/m/s)

Measured Measured
0.1 0.1

0.05 0.05

x (m)
0 0
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 x (m) 0.8

G115J60: u0 = 0.59 m/s, C0 = 1.10 g/L, G115J40: u0 = 0.39 m/s, C0 = 1.91 g/L,
d50 = 115m, ws = 1.01 cm/s d50 = 115m, ws = 1.04 cm/s

Figure 5.19: Comparison of measured and particle tracking predicted deposition


rate profiles of experiments using spherical glass particles (Lee, 2010).

114
0.3 0.3
IP3J80 IP3J70
0.25 0.25
Full Model Full Model
0.2 Simp. Model Simp. Model
0.2
Fs (g/m/s)

Fs (g/m/s)
Measured Measured
0.15 0.15

0.1 0.1

0.05 0.05
x (m) x (m)
0 0
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8

IP3J80: u0 = 0.79 m/s, C0 = 1.17 g/L, d50 = 716m, ws = IP3J70: u0 = 0.69 m/s, C0 = 1.45 g/L, d 50 = 716m, ws =
2.02 cm/s 2.01 cm/s

0.3 0.3
IP3J60 IP3J50
0.25 Full Model 0.25 Full Model
Simp. Model Simp. Model
0.2 0.2 Measured
Fs (g/m/s)

Fs (g/m/s)

Measured
0.15 0.15

0.1 0.1

0.05 0.05
x (m) x (m)
0 0
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8

IP3J60: u0 = 0.59 m/s, C0 = 1.90 g/L, d50 = 716m, ws = IP3J50: u0 = 0.49 m/s, C0 = 2.04 g/L, d 50 = 716m, ws =
2.16 cm/s 2.15 cm/s

0.3 0.3
IP3J40 IP3J30
0.25 0.25 Full Model
Full Model
Simp. Model Simp. Model
0.2 0.2
Fs (g/m/s)
Fs (g/m/s)

Measured Measured
0.15 0.15

0.1 0.1

0.05 0.05
x (m) x (m)
0 0
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8

IP3J40: u0 = 0.39 m/s, C0 = 2.63 g/L, d50 = 716m, ws = IP3J30: u0 = 0.29 m/s, C0 = 3.75 g/L, d 50 = 716m, ws =
2.16 cm/s 2.00 cm/s

Figure 5.20: Comparison of measured and particle tracking predicted deposition


rate profiles of experiments using spherical plastic particles (IP3).

115
0.3 0.3
IP3J66 IP3J50
0.25 Full Model 0.25 Full Model
Simplified Model Simplified Model
0.2 0.2
Fs (g/m/s)

Fs (g/m/s)
Measured Measured
0.15 0.15

0.1 0.1

0.05 0.05
x (m) x (m)
0 0
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8

IP3J66: u0 = 0.65 m/s, C0 = 1.56 g/L, IP3J50: u0 = 0.49 m/s, C0 = 2.26 g/L,
d50 = 716m, ws = 2.28 cm/s d50 = 716m, ws = 2.32 cm/s

0.3 0.3
PJ66 PJ50
0.25 Full Model 0.25 Full Model
Simplified Model Simplfied Model
0.2 0.2 Measured
Fs (g/m/s)
Fs (g/m/s)

Measured
0.15 0.15

0.1 0.1

0.05 0.05
x (m) x (m)
0 0
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8

PJ66: u0 = 0.65 m/s, C0 = 1.57 g/L, PJ50: u0 = 0.49 m/s, C0 = 2.09 g/L,
deq = 634m, ws = 2.08 cm/s deq = 634m, ws = 2.08 cm/s

Figure 5.21: Comparison of measured and particle tracking predicted deposition


rate profiles of experiments using plastic particles (Lee, 2010).

116
0.1 0.1
MFJ80 MFJ70
0.08 0.08
Full Model Full Model
Simp. Model Simp. Model
Fs (g/m/s)

Fs (g/m/s)
0.06 0.06 Measured
Measured

0.04 0.04

0.02 0.02
x (m) x (m)
0 0
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8

MFJ80: u0 = 0.79 m/s, C0 = 0.47 g/L, d50 = 347m, ws = MFJ70: u0 = 0.69 m/s, C0 = 0.48 g/L, d50 = 347m, ws =
2.18 cm/s 2.33 cm/s

0.1 0.1
MFJ60 MFJ50
0.08 0.08
Full Model Full Model
Simp. Model Simp. Model
Fs (g/m/s)

Fs (g/m/s)

0.06 0.06 Measured


Measured

0.04 0.04

0.02 0.02
x (m) x (m)
0 0
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8

MFJ60: u0 = 0.59 m/s, C0 = 0.69 g/L, d50 = 347m, ws = MFJ50: u0 = 0.49 m/s, C0 = 0.82 g/L, d50 = 347m, ws =
2.22 cm/s 2.22 cm/s

0.1
MFJ40
0.08
Full Model
Simp. Model
Fs (g/m/s)

0.06
Measured

0.04

0.02
x (m)
0
0 0.2 0.4 0.6 0.8

MFJ40: u0 = 0.39 m/s, C0 = 0.80 g/L, d50 = 347m, ws =


2.20 cm/s

Figure 5.22: Comparison of measured and particle tracking predicted deposition


rate profiles of experiments using melamine particles.

117
5.6.2 2D sediment deposition profiles
Fig. 5.23 shows the comparison of the observed and predicted sediment deposition
profiles of one experiment, CJ58. The bottom deposition pattern are predicted
with favourable agreement by the particle tracking model for both longitudinal
and transverse extent. It is noticed that in the observation, the deposition is not
symmetric about the longitudinal axis, but tends to deposit more on the lower
side of the axis and shows a slight curving pattern.
For the 2D deposition profiles measured in the present experiments (Glass
particles: Figs. 5.24-5.26), the predicted extent of the deposition compares rea-
sonably well with the observations. The performance of the full model and
simplified model are very similar. Similar to Fig. 5.23, the deposition profile
is not symmetric in the transverse direction, except for the cases with low jet
velocity and high settling velocity (e.g. G215J60 and G180J50). The most sig-
nificant asymmetric profile is the one of G115J60 with high flow velocity and
smallest settling velocity (Fig. 5.26a). This experiment is repeated for several
times and it is observed that such asymmetric bottom deposition is not repeat-
able as the asymmetry in external current is generated at random. The closer
to the jet nozzle, the greater is the asymmetry in the profile. In addition, the
predicted transverse spread is smaller than observation, leaving a higher peak at
the centerline.
Previous cross-section particle concentration measurement (Lee, 2010, also
see e.g. Fig. 5.34) shows that as the particles fall out from the turbulent re-
gion of the jet, they are not falling vertically as expected, but with an inclined
trajectory. The particle trajectories are not stable, but tends to swing across
the cross-section periodically. The instability of the external flow is probably
due to the limited size of the experimental tank and the interaction with the
settling particles. The closer to the jet nozzle (the greater the entrainment flow),
the more prominent is this phenomenon. This is probably the reason for the
asymmetry and the increased spread of the observed deposition profiles. In the
particle tracking model, such complex flow phenomenon has not been modelled.
2D deposition profiles of plastic IP3 and melamine particles are shown in
Fig. 5.27 and Fig. 5.28. Similar to the previous observations, the extent of
the deposition depends on the jet flow velocity. Asymmetry is observed in the
transverse direction. In these cases, the full model predicts better agreement
with the experimental observation, while the simplified model usually predicts
narrower and shorter deposition profiles, with a much higher peak deposition.
The result is consistent with the conclusion in 1D modelling.

5.6.3 Importance of different forces in the equation of


motion to the prediction of deposition profile
In the equation of motion, the gravitational and drag forces appears to be the
most important factor governing the motion of sediment particles. The other
forces (fluid acceleration, added mass and Basset) is of different importance in

118
a) observation
8
y (cm)
6 CJ58
4

-2

-4

-6
x (cm)
-8
0 5 10 15 20 25 30

b) predicted

Figure 5.23: Observed and predicted (particle tracking) deposition pattern of


the experiment CJ58 (Li, 2006), deq = 166m, u0 = 0.57m/s, ws = 1.97m/s

119
is located at x = 0, y = 0.
G215 particles, d50 = 215m, ws = 2.65m/s. Contour in g/m2 /s. The jet nozzle
Figure 5.24: Observed and predicted deposition profile of the experiment with

0.06 0.06
G215J60 G215J90
Observed Observed
0.04 0.04
1 3 1
0.02 2 0.02 6
6
y (m)

y (m)
2

3
0 0
10

6
-0.02 -0.02 3

1
6 1
2
-0.04 -0.04

-0.06 -0.06
(b) G215J60-2D, u0 = 0.58m/s

(a) G215J90-2D, u0 = 0.88m/s


0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
x (m) x (m)

0.06 0.06
G215J60 G215J90
Predicted Predicted
0.04 0.04
Simplified Model Simplified Model
2 1
0.02 0.02
6
10
120

y (m)

y (m)
10 18 36 3
0 22 0 14
14 1 6
2
-0.02 2 -0.02 1

-0.04 -0.04

-0.06 -0.06
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
x (m) x (m)

0.06 0.06
G215J60 G215J90
Predicted Predicted
0.04 0.04
Full Model 1 Full Model
2
0.02 0.02 3
y (m)

y (m)
0 22 0 6 14
610 14 6 13 10

1
10
2

6
-0.02 2 -0.02
3
-0.04 -0.04

-0.06 -0.06
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
x (m) x (m)
is located at x = 0, y = 0.
G180 particles, d50 = 180m , ws = 1.98m/s. Contour in g/m2 /s. The jet nozzle
Figure 5.25: Observed and predicted deposition profile of the experiment with

0.06 0.06
G180J50 G180J80
Observed Observed
0.04 0.04
2 2 1
0.02 0.02
6 2 4 4
y (m)

y (m)
0 0

1
2 3
6

5
-0.02 2 -0.02 2
1

3
2 4
-0.04 -0.04
1
-0.06 -0.06
(b) G180J50-2D, u0 = 0.49m/s

(a) G180J80-2D, u0 = 0.78m/s


0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5 0.6
x (m) x (m)

0.06 0.06
G180J50 G180J80
Predicted Predicted
0.04 0.04
Simplified Model Simplified Model
0.02 0.02 3 4 1
8 6
121

y (m)

y (m)
2104 12 10

2
0 1 22 2 0
6 18 6 812 3
10 463
-0.02 -0.02 4
1
-0.04 -0.04

-0.06 -0.06
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5 0.6
x (m) x (m)

0.06 0.06
G180J50 G180J80
Predicted Predicted
0.04 0.04 1
Full Model Full Model
2 2 3
0.02 0.02
10 6 4
y (m)

y (m)
6 18 22 1102
0 0
14 14

1
6 3 2 8
2

-0.02 2 -0.02 3

-0.04 -0.04

-0.06 -0.06
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5 0.6
x (m) x (m)
is located at x = 0, y = 0.
G115 particles, d50 = 115m , ws = 1.00m/s. Contour in g/m2 /s. The jet nozzle
Figure 5.26: Observed and predicted deposition profile of the experiment with

0.06
G115J40 0.1
G115J60
Observed Observed
0.04
0.05

0.3
0.02 1
0. 0.3
0.1

0.1
0.5
y (m)

y (m)
0 0.5 0
0.7

0.9
0.7
0.1 0.2
0.3 0.3
-0.02 5 0.3

0.1
0.

0.3
-0.05 0.5
0.1 0.3
-0.04 0.4 0.1
0.1 0.2
-0.1
-0.06
(b) G115J40-2D, u0 = 0.39m/s

(a) G115J60-2D, u0 = 0.59m/s


0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
x (m) x (m)

0.06
G115J40 0.1
G115J60
Predicted Predicted
0.04
0.1 Simplified Model Simplified Model
0.3 0.05
0.02 0.5 0.1
0.7 0.5
122

y (m)

y (m)
1.1 0.9
1.9 00.3.7 1.1
0
000..10.35.9 1.25.3 0 0.9
0.3
0.7 0.
1
1.5 0.1 0.1
-0.02 0.5 0.3 -0.05
-0.04
-0.1
-0.06
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
x (m) x (m)

0.06
G115J40 0.1
G115J60
Predicted Predicted
0.04
Full Model Full Model
0.1 0.3 0.
1 0.05
0.02 0.9 0.1
1.1 0. 0.3
7
y (m)

y (m)
0.5
0.

00.2.7 0.5

0.

1
5

0 0

0.
00.3.7 2.3 1.5

2
1.5 0.9
0.1 0.1.91 0.3
-0.02 0.5 0.3 0.1
0.1
-0.05
-0.04
-0.1
-0.06
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
x (m) x (m)
located at x = 0, y = 0.
IP3 particles, d50 = 716m , ws = 2.2m/s. Contour in g/m2 /s. The jet nozzle is
Figure 5.27: Observed and predicted deposition profile of the experiment with

0.06 0.06
IP3J40 IP3J80
Observed Observed
0.04 0.04 0.25
0.5
0.02 0.5 0.02 0.5
0.75
01.5 1.5
2.5 1
y (m)

y (m)
5
0 3.5 5 0 0.7 1.5

0.25
0.2

2.5
4
5
3 1 0.5 1 1.25
-0.02 2 1.5 -0.02
0.5
0.75 0.5
-0.04 -0.04
0.25
-0.06 -0.06
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5 0.6
(b) IP3J40-2D, u0 = 0.39m/s

(a) IP3J80-2D, u0 = 0.78m/s


x (m) x (m)

0.06 0.06
IP3J40 IP3J80
Predicted Predicted
0.04 0.04
Simplified Model Simplified Model
0.02 0.02 0.25 0.5
1 0.7 0.
.5 2.5 1.21
123

0.5 2.25 5 1 4 25
y (m)

y (m)
7 5 5
0
1.5 8 0 00.711.72 3
6 5 3 .55 1
4 1.5 0.5 .25
0.2 2 1.75
1 1.5
-0.02 -0.02 5
0.25 0.5
-0.04 -0.04

-0.06 -0.06
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5 0.6
x (m) x (m)

0.06 0.06
IP3J40 IP3J80
Predicted Predicted
0.04 0.04 0.25
Full Model Full Model
0.02 0.02 0.5
1 1.5 2 1.5 1
5 0.75

0.
0.

5
y (m)

y (m)
4 5 4.5 0.02.15 11.2.
3 5

75
5
0.

0 0
2.

3.5 75

0.25
1
5

3 2.5 5
1.5 1.5 1.2
-0.02 -0.02
0.2 0.5
5
-0.04 -0.04

-0.06 -0.06
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5 0.6
x (m) x (m)
located at x = 0, y = 0.
MF particles, deq = 347m , ws = 2.2m/s. Contour in g/m2 /s. The jet nozzle is
Figure 5.28: Observed and predicted deposition profile of the experiment with

0.06 0.06
MFJ60 MFJ80
0.1 Observed Observed
0.04 0.04
0.3 0.1

0.1
0.3
0.5
0.7
0.02 0.02
0.7 0.3
1

0.1
0.3
0.
y (m)

y (m)
0 0.5 0
0.7

0.1
0.3 0.5
-0.02 0.1 -0.02
0.3
-0.04 -0.04 0.1
-0.06 -0.06
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5 0.6
(b) MFJ40-2D, u0 = 0.59m/s

(a) MFJ80-2D, u0 = 0.78m/s


x (m) x (m)

0.06 0.06
MFJ60 MFJ80
Predicted Predicted
0.04 0.04
Simplified Model Simplified Model
0.1 0.1
0.02 0.02
0.3 0.3
0.5
1.5
124

0.9 0.7
y (m)

001..57.1.1

y (m)
1.7 0 1.1 1.5
1.7
1.9 1.3
0 0 0.1 .9
00..13 2 2.5
1.1 0.9 0.3
0.7 0.5 0.5
-0.02 -0.02 0.1
0.1
-0.04 -0.04

-0.06 -0.06
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5 0.6
x (m) x (m)

0.06 0.06
MFJ60 MFJ80
Predicted Predicted
0.04 0.04
Full Model Full Model
0.3 0.1
0.02 0.1 0.02
35
0.0. 0 0.1
0.1 0.9.7 0. 0.
y (m)

y (m)
1.5 3 0.7 5
0 1.1 0 0.5 1.3 1.1
1.1
0.7 0.5 0.3
-0.02 -0.02 0.3 1
0.1 0.1 0.1 0.
-0.04 -0.04

-0.06 -0.06
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5 0.6
x (m) x (m)
unsteady turbulent motion. To investigate the importance of these forces, sensi-
tivity study has been carried out in predicting the 1D and 2D deposition profiles.
By excluding the forces one by one in the order of Basset, added mass and fluid
acceleration from the equation of motion, the predicted bottom deposition pro-
files are compared with each other to show their relative importance. The test
on six experimental cases of Qj = 60L/hr (Qj = 58L/hr for sand particle) using
sand, glass and plastic particles are presented.
It is observed that the removal of Basset, added mass and fluid acceleration
forces does not have a significant impact on the predicted bottom deposition
profile, both in terms of 1D (Fig. 5.29) and 2D (Fig. 5.30) profiles. These tran-
sient forces are unimportant to the prediction of particle motion in turbulent
jets. The major cause of the difference in deposition pattern between glass and
plastic particles are due to their inertia. The higher inertia of the plastic par-
ticles results in slower response to turbulent fluctuations and thus increases the
instantaneous drag force and longer suspension time in turbulence. Indeed a very
simple equation of motion containing the buoyancy and drag force terms suffices
for the prediction. In this study, the computationally demanding Basset force
can be excluded for the use in sediment-laden jets, while the other transient force
terms are retained as they pose no difficulty in the numerical solution.

125
0.7 0.7
G215J60 G180J60
0.6 G+D+F+A+B 0.6
G+D+F+A+B
G+D+F+A 0.5
0.5 G+D+F+A
G+D+F
Fs (g/m/s)

Fs (g/m/s)
G+D G+D+F
0.4 0.4
Measured G+D
0.3 0.3 Measured
0.2 0.2

0.1 0.1
x (m)
0 0
0 0.2 0.4 0.6 x (m) 0.8 0 0.2 0.4 0.6 0.8

(a) Glass, G215J60: u0 = 0.59 m/s, C0 = 5.51 g/L, (b) Glass, G180J60: u 0 = 0.57 m/s, C0 = 6.18 g/L,
d 50 = 215m, ws = 2.74 cm/s d50 = 180m, ws = 1.82 cm/s

0.2 0.5
G115J60 FJ58
G+D+F+A+B G+D+F+A+B
0.4
0.15 G+D+F+A
G+D+F+A
G+D+F
Fs (g/m/s)

Fs (g/m/s)

G+D 0.3 G+D+F


0.1 Measured G+D
0.2 Measured

0.05
0.1
x (m) x (m)
0 0
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8

(c) Glass, G115J60: u0 = 0.59 m/s, C0 = 1.10 g/L, (d) Sand, FJ58: u 0 = 0.57 m/s, C0 = 3.93 g/L,
d 50 = 115m, ws = 1.01 cm/s deq = 135m, ws = 1.41 cm/s

0.3 0.1
IP3J60 MFJ60
0.25 G+D+F+A+B 0.08 G+D+F+A+B
G+D+F+A G+D+F+A
0.2 G+D+F G+D+F
Fs (g/m/s)
Fs (g/m/s)

G+D 0.06
G+D
0.15 Measured
Measured
0.04
0.1

0.05 0.02
x (m) x (m)
0 0
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8

(e) Plastic, IP3J60: u0 = 0.59 m/s, C0 = 1.90 g/L, (f) Plastic, MFJ60: u0 = 0.59 m/s, C0 = 0.69 g/L,
d 50 = 716m, ws = 2.16 cm/s deq = 347m, ws = 2.20 cm/s

Figure 5.29: Sensitivity tests on the Basset, added mass and fluid acceleration
forces. 1D deposition profiles. G: gravity; D: drag force; F: fluid acceleration; A:
added mass; B: Basset. (a)-(c): Glass (Lee, 2010); (d): Sand (Li, 2006); (e)-(f):
plastic, present study

126
y = 0.
deposition profiles. Contour in g/m2 /s. The jet nozzle is located at x = 0,
Figure 5.30: Sensitivity tests on the Basset force (present experiments). 2D

0.06 0.06
(a) IP3J60 G180J80
Observed Observed
0.04 0.04
1 0.5
1.5 1
1.5 2

5
0.02 0.02

1 0.
2

2.5

2
y (m)

y (m)
1 6
0 1.5 0
1 5 4 3
0.5 0.5
-0.02 -0.02 1
4 2

2
3
-0.04 -0.04
1
-0.06 -0.06

(a) G180J80-2D, u0 = 0.78m/s


0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5 0.6
(b) IP3J60-2D, u0 = 0.59m/s

x (m) x (m)

0.06 0.06
(b) IP3J60 G180J80
Predicted Predicted
0.04 0.04
w/ Basset force 1 w/ Basset force
0.02 1 0.5 0.02
1.5 3
4
127

6
0.5
y (m)

y (m)

2
2.5 2 12
0 0 8 10
3.5
3 8

1
4
3
1 2 6
-0.02 1.5 1 -0.02 2
0.5 1
-0.04 -0.04

-0.06 -0.06
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5 0.6
x (m) x (m)

0.06 0.06
(c) IP3J60 G180J80
Predicted Predicted
0.04 0.04
w/o Basset force w/o Basset force
0.5 1 1
0.02 0.02 3 2
4
6 8
y (m)

y (m)
12.5.35 4

1
0 2 0
3.5 2
1.
5 23 10
0.5

0. 1 2.5 4
1

-0.02 5 -0.02 1
3 2
-0.04 -0.04 1

-0.06 -0.06
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5 0.6
x (m) x (m)
5.6.4 Cross-sectional sediment concentration profiles
Sediment concentration measured in the jet cross-section using particle imaging
technique (Lee, 2010) is utilized as comparison with the predicted concentration
profiles. Transformation of particle mass to sediment concentration is required for
the discrete particle tracking model. Within a control volume V = xyz,
the concentration can be evaluated with the average number of particles inside it.
The transverse grid size (y and z) is defined as one sixth of the top-hat width,
similar to the grid size used in particle imaging measurement. The thickness of
the grid x, is half the diameter of the jet = 3mm.
Figs. 5.31 to 5.34 show the comparison of measured and particle tracking
predicted cross-sectional sediment concentration profiles of four of the experi-
mental cases by Lee (2010). The results presented here are all obtained by the
simplified model. For all the cases, the sediment concentration profiles in three
longitudinal locations, respectively, x < lm , x lm and x > lm , are included.
The top-hat half-width of the jet, defined as bT = 0.161x, is included for the ease
of comparison.
The particle concentration across the sediment momentum jet is in general
horseshoe-shaped, except for the case of G115J70 (Fig. 5.34, the highest lm among
all cases) with the location close to the jet nozzle, where the settling of particles
is counter-balanced by jet turbulence and the entrainment flow. For x < lm ,
the maximum concentration (enclosed by the contour of 0.8Cmax ) is well defined
inside the jet top-hat turbulent region. Most of the contour line are still enclosed,
though the outer ones become elliptically elongated downwards. The vertical
location maximum particle concentration Cm is approximately at the center of
the jet, while some of the sediment start to settle out from the jet. For x lm ,
the sediment cloud starts to depart from the water jet; the location of maximum
concentration starts to drop downwards, but still within the top-hat jet boundary.
The contour of 0.8Cmax is still enclosed but the concentration pattern is much
elongated in vertical direction and becomes elliptic. For x > lm , the particle cloud
separates significantly from the water jet; a complete horseshoe profile appears
as all the concentration contours are no longer closed. The maximum particle
concentration (actually the line of the maximum concentration) is located below
the top-hat jet region. In the measurement, the horseshoe trail is often skewed
to one side reflecting the complexity and instability of the external entrainment
flow.
The predicted cross-sectional concentration contours in general compares very
well to the experimental measurements. For some cases the predicted concen-
tration width (defined by 0.25Cmax is somewhat narrower than the observation.
However, given the measurement error of the particle concentration by the imag-
ing method (the identification of a particle depends on a threshold light intensity
which has to be calibrated using the result from other measurement methods,
e.g. suction), the discrepancy is considered to be minor and does not affect the
subsequent discussions.
Figs. 5.35 and 5.36 are the comparison of measured and predicted cross-
section concentration profiles for plastic IP3 particles in the present experiments

128
using the full model. It is observed that, due to the particle inertia, plastic parti-
cles are maintained in the jet for a longer distance by jet turbulence. For example,
in the case of IP3J80, x/D = 12.5 < lm /D (Fig. 5.35), particles have not yet
significantly fallen out from the jet. The concentration contours remain nearly
circular like a pure jet. On the other hand for the glass particle case of G180J80
(Fig. 5.31), significant sediment fall out can be observed at the same location
(x/D = 12.5) as the outer concentration contours become elongated ellipses.
The predicted cross-sectional concentration profiles of IP3 particles compares
very well with the experimental measurements.
The comparison of predicted and measured maximum sediment concentra-
tion at the centerline plane is shown in Figs. 5.37-5.39. The measurement and
predicted concentrations compares very well. The result from the full model on
the glass particles are very similar to the simplified model with less than 5%
difference in the maximum concentration for each plane. Such difference would
already been resulted from the nature of the stochastic model. When x < lm , the
sediment concentration lies very close to the theoretical centerline concentration
profile for a simple jet with neutral buoyant tracers. As x > lm , the center-
line maximum concentration departs rapidly from the theoretical values due to
the massive fall out of sediment particles. Fig. 5.40 shows the predicted and
measured centerline maximum sediment concentration for plastic IP3 particles.
The full model shows better performance over the simplified model on plastic
particles, especially on x > lm .

129
Measured Predicted
4 4

3 3

2 2
0. 25
0 .4
1 0.6
1

0. 25

0.
0.2
0.06.8

25 0.4
4
z/D

z/D

0.
0 0

5
0.25

0.6
0.4

-1 -1
0.8

0.4
-2 -2
6

5
0.

0.2

0.25

0 .2 5
0.
4
-3 -3
5
0.2

0.4

-4 -4
-4 -3 -2 -1 0 1 2 3 4 -4 -3 -2 -1 0 1 2 3 4
y/D y/D

x/D = 12.5 (x < lm)

8 8

6 6

4 4

2 0. 2
25
0.4 0.25
0 .6
z/D

z/D

0 0 0.6
5
0.2

0.

0.4
0.8
8

-2 -2
0.4

0.25
0.4
0.2 5
0.4

0.25
0 .6

-4 -4
0.6
0.6
0.
8

0.6

-6 -6
0 .25

0.4

-8 -8
-5 0 5 -5 0 5
y/D y/D

x/D = 25.7 (x lm)

10 10

5 5
z/D

z/D

0 0 5
0 .4 0.2
0.2
5
0.2

0.6
5

0 .8
0.2 5
0.4

0.6
0.4

-5 -5
0.4

0.6

8
0.
0.8
0.4

0.6

0.25

0.8

-10 -10
5
0.2
5
0.6
0.2

0.8

-10 -5 0 5 10 -10 -5 0 5 10
y/D y/D

x/D = 41.7 (x > lm)

Figure 5.31: Measured and particle tracking predicted (simplified model) cross-
sectional sediment concentration (normalized with the maximum), Case G180J80
(Lee, 2010), u0 = 0.78m/s, ws = 2.06cm/s, lm /D = 32.9. Dashed circle repre-
sents the top-hat profile of the jet.

130
Measured Predicted
4 4

3 3

2 2

1 0.4
1
0.4
25

0.
25 0.6

25
0.
0.
z/D

z/D
0 0

0 .25
0.6

0.8
-1 -1

0.4
0.6
0.8
0.4

0.8

0.4
-2 -2

0.6
0.25
0.2

0.4
5

0.25
-3 -3
0.6

0.25
-4 -4
-4 -3 -2 -1 0 1 2 3 4 -4 -3 -2 -1 0 1 2 3 4
y/D y/D

x/D = 12.5 (x < lm)

6 6

4 4

2 2

25 0.4
0. 2
5 0
0.
z/D

z/D

.4
0 0
0 .2 5

.6
0. .6

0
0
8

-2 -2 0.25
0.8
0.6
0.4
0.4

0.8

0.25
0.4

0.4
0.6

-4 -4
0.25

0.6
0.25
0.6

-6 -6
-6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6
y/D y/D

x/D = 20.8 (x lm)

10 10

5 5
z/D

z/D

0 0 0.2
5
5
0.2

0.25

0.6
0 .8
0.4

0.4

0.4
0.4

-5 -5
0.6

0.25
0.6

0.8

0.25
0.6
0.25

0.8

-10 -10
-10 -5 0 5 10 -10 -5 0 5 10
y/D y/D

x/D = 33.3 (x > lm)

Figure 5.32: Measured and particle tracking predicted (simplified model) cross-
sectional sediment concentration (normalized with the maximum), Case G180J50
(Lee, 2010), u0 = 0.49m/s, ws = 2.05cm/s, lm /D = 20.8. Dashed circle repre-
sents the top-hat profile of the jet.

131
Measured Predicted
4 4

3 3

2 2

1 1
0.25
25 0. 4
0.
z/D

z/D
0 0 0.6

0.4
0.6

0.25

0 .8
-1 -1

0.25
0.4
0.8
0.4

0.25
0.8

-2 -2

0.6
0.25

0.6
0.4

0.6

0.4
-3 -3
5
0.6

0.2

-4 -4
-4 -2 0 2 4 -4 -2 0 2 4
y/D y/D

x/D = 12.5 (x < lm)

6 6

4 4

2 2

5
0.2
z/D

z/D

0 0 .25 0
0.
8 6
0.
0.4

0.6 0 .4
0.4

-2 -2
0.25
0.4

0.25
0.8

0.6

-4 -4
0.25

0.8

0.25

0. 8
0.6

0.4
0.6

0.8

0.6

-6 -6
-5 0 5 -5 0 5
y/D y/D

x/D = 20.8 (x lm)

10 10

5 5
z/D

z/D

0 0
5 0.2
0.2 5
0.4

0 .6
0.8
0.25
0.4
0.4

-5 -5
0.6

0.25

0.4
0 .8
0.25

0.6
0.6

0.25
0.8

-10 -10
-10 -5 0 5 10 -10 -5 0 5 10
y/D y/D

x/D = 33.3 (x > lm)

Figure 5.33: Measured and particle tracking predicted (simplified model) cross-
sectional sediment concentration (normalized with the maximum), Case G215J60
(Lee, 2010), u0 = 0.59m/s, ws = 2.69cm/s, lm /D = 18.9. Dashed circle repre-
sents the top-hat profile of the jet.

132
Measured Predicted
4 4

3 3
0.3
2 0.4 2

3
6 0.6
0.
1 0. 1
0.4 0.4

0.2
5
0.60 .8

0 .2
z/D

z/D

5
0 0
0 .8

0.3
0.4
-1 -1

0 .4
0.6
0.3

-2 -2

0.
25

4
0.4
0.

0.
2
-3 -3

5
3
0.

-4 -4
-4 -3 -2 -1 0 1 2 3 4 -4 -3 -2 -1 0 1 2 3 4
y/D y/D

x/D = 12.5 (x < lm)

15 15

10 10

5 5
5
25

0.4 0.2
z/D

z/D
0.

0 0 4
0.
0.25
0 .6

-5 -5
0.6

0.25
0.8

0.4
0.6
0.8

0.25
0.6
0.4

-10 -10
0.25

0.8

.4

0.4
0.2 0

-15 -15
5
0.6

-15 -10 -5 0 5 10 15 -15 -10 -5 0 5 10 15


y/D y/D

x/D = 58.3 (x lm)

20 20

15 15

10 10

5 5
0.2
5
z/D

z/D

0 0
5 0. 4
0.2 0 .8
-5 -5
0.25
0.4

0 .6
0.4
0.6
0.25

-10 -10
0.25
4
25
0.

0.4

-15 -15
0.8
0.

0.6

0.
8

-20 .6 -20
-20 -15 -10 -5 0 5 10 15 20 -20 -15 -10 -5 0 5 10 15 20
y/D y/D

x/D = 66.7 (x > lm)

Figure 5.34: Measured and particle tracking predicted (simplified model) cross-
sectional sediment concentration (normalized with the maximum), Case G115J70
(Lee, 2010), u0 = 0.68m/s, ws = 1.03cm/s, lm /D = 57.6. Dashed circle repre-
sents the top-hat profile of the jet.

133
Measured Predicted
4 4

3 3

2 5
2
0.2
1 1 25
0. 0.4
0.

0 .4
0.25
0.6
0.4 6

0.2
8
0.
z/D

z/D
0 0

5
8
0.4
0.6

0.
0 .25

0.6
-1 -1
0.4

0.4
0.4

0.2
-2 2 5 -2
0.

0.25
-3 -3

-4 -4
-4 -3 -2 -1 0 1 2 3 4 -4 -3 -2 -1 0 1 2 3 4
y/D y/D

x/D = 12.5 (x < lm)

10 10

5 5
0.25

0.6
0.25 0
0 .4
0.4

.4
z/D

z/D

0 0 0.6
0.25

0.8
0.8

0.25
0.25
0.8

0.4
0.6

-5 -5
0.25
0.6

0.
0.4
0.6

4
0.6
0.4

0.
25
0 .6

-10 -10
-10 -5 0 5 10 -10 -5 0 5 10
y/D y/D

x/D = 33.3 (x lm)

15 15

10 10

5 5
0.2 5
0.4 0.6
z/D

z/D

0 0
0.4 0.0.6
0.2

8
5
0.25

0.25

-5 -5
5
0.8

0 .6 0.8

0.2.4
0

-10 -10
0.4

0.6

0.4
0 .6

0.25
0.8
0.8
0.4

0.6

-15 0.8 -15


0.2

-15 -10 -5 0 5 10 15 -15 -10 -5 0 5 10 15


y/D y/D

x/D = 50.0 (x > lm)

Figure 5.35: Measured and particle tracking predicted (full model) cross-sectional
sediment concentration (normalized with the maximum), Case IP3J80 (present
experiment), u0 = 0.78m/s, ws = 2.02cm/s, lm /D = 34.5. Dashed circle repre-
sents the top-hat profile of the jet.

134
Measured Predicted
4 4

3 3

2 2

1 0.4 1

0. 2

0. .4
25
5 6

5
0. 0.8
2

25
0 .6
z/D

0.
0.

z/D
0

0
0

0.4

0.8
-1 -1 0.6

0.2 0.4
0.6

5
0.4

-2 -2

0.2
0 .4

0.25
5
0.
0.2

-3 -3
5

-4 -4
-4 -3 -2 -1 0 1 2 3 4 -4 -3 -2 -1 0 1 2 3 4
y/D y/D

x/D = 12.5 (x < lm)


6 6

4 4

2 0.25 2
0.4
0.4 0.2
0.6
z/D

y/D

5
0 0
0.2 5

0.25
0.4

0.6
0.

8
0.8
0.6
0.8

0.4
0.2 5

-2 -2
0.
8

0 .6

0.25
0.4
0.6

0.4

-4 -4
0.25
0.25

0.4

0.6

-6 -6
-6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6
y/D x/D

x/D = 20.8 (x lm)

10 10

5 5

5
0.2
z/D

z/D

0 0 5
0.
4 0.2
6

0 .6
0.4
0.

0.2
0.8

0.8

0.4
0 .4
5

-5 -5
0.25

0.25
0.8
0.6

0.25

0.6
0.8
0.6
0.8

0.6
0.4
0.4

-10 -10
-10 -5 0 5 10 -10 -5 0 5 10
y/D y/D

x/D = 33.3 (x > lm)

Figure 5.36: Measured and particle tracking predicted (full model) cross-sectional
sediment concentration (normalized with the maximum), Case IP3J50 (present
experiment), u0 = 0.49m/s, ws = 2.02cm/s, lm /D = 21.5. Dashed circle repre-
sents the top-hat profile of the jet.

135
10 10
Cmax /C0 Cmax/C0
Meas. Meas.
Full Model
Full Model
Simp. Model
Simp. Model Free jet
1 1
Free jet

0.1 0.1

lm /D = 28.3 lm/D = 18.9


0.01 0.01
x/D x/D
1 10 100 1000 1 10 100 1000

(a) G215J90: u0 = 0.88m/s, lm /D = 28.3 (b) G215J60: u0 = 0.59m/s, lm /D = 18.9

Figure 5.37: Measured and particle tracking predicted centerline maximum con-
centration, G215 particles, ws = 2.69cm/s (Lee, 2010). Dashed line is the theo-
retical tracer concentration variation for a free jet.

10 10
Cmax/C0 Cmax/C0
Meas. Meas.
Full Model Full Model
Simp. Model Simp. Model
1 Free jet 1 Free jet

0.1 0.1

lm/D = 32.9 lm/D = 20.6


0.01 0.01
x/D x/D
1 10 100 1000 1 10 100 1000

(a) G180J80: u0 = 0.78m/s, lm /D = 32.9 (b) G180J50: u0 = 0.49m/s, lm /D = 20.6

Figure 5.38: Measured and particle tracking predicted centerline maximum con-
centration, G180 particles, ws = 2.06cm/s (Lee, 2010). Dashed line is the theo-
retical tracer concentration variation for a free jet.

136
10 10
Cmax /C0 Meas. Cmax /C0
Meas.
Full Model Full Model
Simp. Model
Simp. Model
Free jet
1 1 Free jet

0.1 0.1

lm /D = 57.6 lm/D = 32.9


0.01 0.01
x/D x/D
1 10 100 1000 1 10 100 1000
(a) G115J70: u0 = 0.68m/s, lm /D = 57.6 (b) G115J40: u0 = 0.39m/s, lm /D = 32.9

Figure 5.39: Measured and particle tracking predicted centerline maximum con-
centration, G115 particles, ws = 1.03cm/s (Lee, 2010). Dashed line is the theo-
retical tracer concentration variation for a free jet.

10 10
Cmax/C0 Cmax/C0
Meas. Meas.
Full Model Full Model
Simp. Model Simp. Model
1 Free jet 1 Free jet

0.1 0.1

lm/D = 34.5 lm/D = 21.5


0.01 0.01
x/D x/D
1 10 100 1000 1 10 100 1000

(a) IP3J80: u0 = 0.78m/s, lm /D = 34.5 (b) IP3J50: u0 = 0.49m/s, lm /D = 21.5

Figure 5.40: Measured and particle tracking predicted centerline maximum con-
centration, IP3 particles, ws = 2.02cm/s (present experiments). Dashed line is
the theoretical tracer concentration variation for a free jet.

137
5.7 Summary
In this chapter, the mixing and deposition of sediment-laden horizontal mo-
mentum jets are studied using laboratory experiments and computational fluid
dynamics modelling (CFD). A stochastic particle tracking model is developed to
model the bottom deposition and sediment concentration in horizontal sediment-
laden momentum jets.
Experimental flow visualization shows that the jet fluid phase is not sig-
nificantly altered by the presence of sediment under the dilute concentration
(C0 < 7.7g/L, volume fraction < 0.3%) in this study. Dimensional analysis of
experimental measurement shows that for the same type of particles, the 1D
bottom deposition profiles collapse into a single log-normal profile upon normal-
ization by the jet momentum-settling length scale lm and the source sediment
mass flux.
CFD modelling shows that there is a significant settling velocity reduction
up to about 25-35% under the influence of jet turbulence, which is dependent
on the RMS turbulent velocity to stillwater settling velocity ratio /ws and the
particle properties. Using a relationship for estimating the reduction in settling
velocity locally based on the /ws ratio, the Eulerian CFD model predicted
bottom deposition of sand and glass particle-laden jets are significantly improved
as compared to using directly the stillwater settling velocity. Alternatively, a
whole field reduction of 25% on the stillwater settling velocity works well for
sand and glass particles, nevertheless a 35% whole-field reduction is required for
plastic particles.
The stochastic particle tracking model adopts the autocorrelation function
developed in Chapter 3 which accounts for the trapping and loitering effect of
particles in turbulent eddies. Analytical axial and radial mean flow solutions
for a pure jet is used. Turbulent velocity fluctuations and turbulent time scales
are modelled with best-fitted self-similar RMS turbulent velocity and dissipation
rate profiles derived from CFD solution of a pure jet. The particle tracking
model is applied to predict the deposition and concentration profiles of horizontal
sediment-laden jets. The model has been fully tested on existing sediment-laden
jet experimental data, with initial jet velocity u0 = 0.3 0.9m/s and covering a
wide range of particle properties, including

Sand (p = 2.65 g/cm3 , d = 133 166m , ws = 1.4 2.0cm/s),

Glass (p = 2.5 g/cm3 , d = 115 215m , ws = 1.0 2.7cm/s),

Plastic(MF) (p = 1.5 g/cm3 , d = 347m , ws = 2.2cm/s), and

Plastic(IP3) (p = 1.14 g/cm3 , d = 716m , ws = 2.0cm/s).

Model predictions are in excellent agreement with experimental data. The parti-
cle tracking model is superior to a 3D Eulerian CFD model as it does not require
any a priori empirical adjustment to the stillwater settling velocity.

138
Comparison with experimental data shows that the simplified particle track-
ing model (assuming particle responses to change in fluid velocity immediately)
works well for sediment-laden jet with sand and glass particles. However, when
lighter plastic particles of relative density 1.1 and 1.5 are used, a more general
model (the Full Model) based on solving the full equation of motion of sedi-
ment particles needs to be used to explain the deposition profiles of horizontal
sediment-laden jets with a wide range of particle density and sizes. Supported by
experimental data, a sensitivity study shows that the computationally demand-
ing Basset history force can be neglected for all kind of particles used in present
sediment-laden jet experiments, thus greatly simplifying the model. The model
is further generalized for arbitrarily inclined buoyant jets in in Chapter 6.

139
140
Chapter 6

A General Model of
Sediment-Laden Buoyant Jets

6.1 Introduction
A particle tracking model has been developed to predict the deposition and
sediment concentration of horizontal sediment-laden momentum jets in Chapter
5. The predictions are in excellent agreement with experimental data. However,
many environmental flows are also influenced by the density difference between
the effluent and the ambient. For example, in submarine sewage discharge the
effluent density is similar to that of freshwater (1000 kg/m3 ), while seawater
has a density of about 1025 kg/m3 (salinity = 33ppt). The sewage outfalls are
usually designed such that the jets are discharged in a horizontal direction to
maximize turbulent mixing. The jet flow is initially horizontal, but eventually
the buoyancy dominates the dynamics of the flow and results in bending up of
the jet. Ambient stratification and jet-induced external flow further complicate
the problem.
In this chapter, a general particle model for predicting the bottom deposition
of sediment-laden buoyant jet in stagnant condition is developed. The model is
based on the theoretical development in Chapter 3 and essentially an extension
from the model for sediment-laden momentum jet in Chapter 5. Nevertheless,
the model incorporates the three flow regimes affecting the sediment dynamics,
namely the turbulent jet flow, external entrainment flow and surface spreading
current. The three mean flow regimes can be determined analytically and/or
semi-analytically at any location. The model prediction of sediment deposition
is validated by extensive experimental measurements in previous studies.

6.2 The flow regimes in buoyant jets


For a buoyant jet, the three flow regimes involved are (Fig. 6.1):

1. the buoyant jet flow characterized by intense turbulence,

141
2. the external irrotational flow field induced by jet entrainment, and

3. the spreading current as the jet impinges the surface.

A sediment particle, once ejected from the jet nozzle, would at least encounter the
first and second flow regime before they deposit on the bottom. Some particles
which are lighter and carried by stronger flows will rise to the water surface and
encounter all the three regimes. Some of the particles that fall out from the jet
or the spreading current may even be re-entrained back into the jet flow before
they eventually deposit on the bottom.

Jet-induced
external flow

Figure 6.1: A general schematic representation of a sediment laden buoyant jet.

6.2.1 The buoyant jet flow


General phenomenon

The buoyant jet flow regime is characterised by the intensive turbulence induced
by shearing with the ambient environment. A single-phase horizontal buoyant
jet is governed by the initial kinematic momentum flux M0 = Q0 u0 and initial
specific buoyancy flux B0 = Q0 g0 . The jet densimetric Froude number
q
F r = u0 / g0 D (6.1)

characterizes the relative importance of momentum and buoyancy of a buoyant


jet. Here u0 is the initial jet velocity; Q0 = u0 D2 /4 is the jet flow rate; g0 =
(/a )g is the reduced gravity; is the density difference between the jet and
ambient fluid; a is the ambient fluid densities; D is the jet diameter. When

142
F r is large, the buoyant jet behaves like a momentum jet for a large part of its
length. On the other hand, the jet becomes a plume quickly when F r is small.
A momentum-buoyancy length scale can be defined as
3/4 1/2
ls = M0 /B0 (6.2)
to describe the transition of momentum dominated region to buoyancy-dominated
region. Close to the jet exit (x << ls ), the buoyant jet is dominated by the initial
momentum and behaves like a pure momentum jet. The jet entrains the am-
bient fluid and gradually rises by the buoyancy force acting vertically upwards.
At a distance x ls , the buoyancy effect becomes comparable to the initial mo-
mentum. In the region x >> ls the buoyant jet is dominated by buoyancy and
behaves like a plume. The entrainment velocity is directly proportional to the
centerline velocity uc and so it varies x1 with the downstream distance in the
jet-like region, and x1/3 in the plume-like region.

Modelling of mean flow

In the present study, the buoyant jet flow is modelled using the integral La-
grangian model JETLAG (Lee and Cheung, 1990; Lee and Chu, 2003) (Fig. 6.2).
JETLAG predicts the mixing of buoyant jets with three-dimensional trajectories
in a wide range of ambient conditions (e.g. stratification, crossflow) using the
entrainment hypothesis. The unknown jet trajectory is viewed as a sequential
series of non-overlapping plume elements which increase in mass as a result of
shear-induced entrainment close to the jet discharge, and vortex entrainment
due to the crossflow. The model adopts a top-hat profile and tracks the average
properties of a plume element at each step by conservation of horizontal and
vertical momentum, conservation of mass accounting for entrainment, and con-
servation of tracer mass/reduced gravity. It has been validated against extensive
basic experimental and field data over a wide range of ambient conditions. For
the present study only the shear entrainment is invoked for stagnant conditions.
The key of the JETLAG model is the entrainment hypothesis for turbulent
closure. For a buoyant jet in stagnant ambient, the increase in jet volume flux
dQ along the jet trajectory s is expressed in terms of the local jet velocity and
width via an entrainment coefficient. It assumes that the local jet radial entrain-
ment velocity ve is proportional to the local streamwise jet velocity, taken as the
maximum velocity uc or an average velocity uT .
dQ/ds = 2bg ve (6.3)
ve = g uc = T uT (6.4)
The entrainment coefficient g or T is found to be different for the asymptotic
jet (g = 0.057) and plume (g = 0.088) regimes. The general dependence of
the entrainment coefficient on the local jet densimetric Froude number can be
derived as (Fox, 1970; Lai and Lee, 2012b)
1 0.554 sin k
g = T = 0.057 + (6.5)
2 Fl2

143
where k is the jet orientation angle with respect to the vertical. Fl is the local
densimetric Froude number defined as
F uT,k
Fl = r (6.6)
T,k
a
gbT,k

where the subscript k denotes the kth jet element and all other values are defined
as top-hat profile values, F is a proportional constant = 1.8. The entrainment
coefficient has the asymptotic values of 0.057 for a pure jet (Fl ) and 0.088
for that of a pure vertical plume (Fl = 4.2). All the quantities (uT , T , bT ) for
computing the local densimetric Froude number can be readily obtained from
the prediction of JETLAG.
For the use in the particle tracking model, the top-hat profile of the JETLAG
prediction has to be converted to the corresponding Gaussian profiles by the
following equations derived from the mass and momentum balance of a jet cross
section (Fig. 6.3).

uc = 2uT (6.7)

bg = 2bT (6.8)
1
g = T (6.9)
2
1 + 2
()c = ()T (6.10)
2
where the subscripts g and T denote the Gaussian and top-hat properties respec-
tively; subscript c denotes the centerline values; u is the streamwise jet velocity;
b is the jet half-width; is the shear entrainment coefficient dependent on lo-
cal densimetric Froude number and jet orientation (Eq. 6.6); is the density
difference between the jet and the ambient; = 1.2 is the ratio between the
Gaussian half-width of concentration and velocity.
Along the jet trajectory, the cross section mean flow are described by a Gaus-
sian profile. The velocity profile in the streamwise direction s is
!
u(s, r) r2
= exp 2 (6.11)
uc (s) bg

The radial velocity normal to the jet trajectory can be derived from the continuity
equation (Lee and Chu, 2003)
Z 
d r

u(s, r )2r dr = 2rvr (r) (6.12)
ds 0

vr (r) [1 exp(r2 /b2g )] (/g )(r2 /b2g ) exp(r2 /b2g )


= (6.13)
uc r/bg
where the jet spreading rate

= dbg /ds. (6.14)

144
Figure 6.2: The JETLAG model (Lee and Cheung, 1990; Lee and Chu, 2003).

u/uc
1

0.8 Gaussian Profile

0.6
Top-Hat Profile
uT = 0.5uc
0.4

0.2
r/bg
0
bT = 2bg
-3 -2 -1 0 1 2 3

Figure 6.3: The Gaussian variation of mean axial velocity and the equivalent
top-hat profile.

145
The original JETLAG model does not distinguish between the potential core
(zone of flow establishment, ZFE) and the fully turbulent self-similar region
(zone of established flow, ZEF). However, testing calculation of a pure momen-
tum jet shows that if the ZFE is excluded (the jet computation starts from the
nozzle), the predicted mean flow velocity is smaller than analytical free jet solu-
tion (Fig. 6.4) which results in lower entrainment flow and turbulence intensity,
causing more sediment to drop out close to the jet nozzle which is incomparable
to the experimental data. In the present study the potential core is separately
determined before the commencement of JETLAG calculation. Unlike a pure
momentum jet which the potential core length is dependent on the jet diame-
ter only (6.2D), the length of potential core of a buoyant jet/plume varies with
the initial buoyancy. The trajectory of the potential core is predicted with the
balance in horizontal and vertical momentum, by assuming the spreading rate
of the turbulence development region to be constant (see Lee and Chu, 2003).
JETLAG computation starts at the end of the potential core with the predicted
top-hat width there and ceases as the upper top-hat boundary hits the water
surface.

Modelling of turbulence

Similar to the modelling of momentum jet in Chapter 5, the stochastic turbu-


lent fluctuation is modelled using the RMS turbulent velocity and turbulent
energy dissipation rate to predict the time and length scales of turbulence,
with Eqs. 3.21 and 3.20. CFD prediction of a vertical buoyant jet (u0 = 1.0m/s,
D = 6mm and F r = 50) shows that the RMS turbulent velocity fluctuations
and turbulent energy dissipation rate profiles in the asymptotic jet and plume
regimes are similar if they are normalized with the centerline velocity uc and
Gaussian jet half-width bg (Fig. 6.5). This is also supported by previous experi-
mental measurement
which showed little difference in the RMS turbulent velocity
fluctuation (e.g. u u ) between a pure jet and a pure plume (Papanicolaou and
List, 1988; Wang and Law, 2002). Thus a single profile of normalized and
are used, using the best fitted equations of Eqs. 5.7 and 5.8, disregarding the
asymptotic jet and plume regimes.

146
uc/u0
10

Zone of flow Zone of


establishment established flow

Theoretical
6.2(D/x)
0.1 JETLAG w/
ZFE
JETLAG
w/o ZFE x/D
0.01
1 10 100
 1
Figure 6.4: Comparison of theoretical ucu(x) 0
= 6.2 Dx and JETLAG model
predicted centerline velocity for a pure jet (u0 = 0.786m/s, D = 6mm), showing
the effect of inclusion of the zone of flow establishment (ZFE).

0.5 0.5
/uc 15D, Fr_l = 15 (b)1/3/uc 15D, Fr_l = 15
0.4 0.4 20D, Fr_l = 12
20D, Fr_l = 12
70D, Fr_l = 5
70D, Fr_l = 5
0.3 0.3 85D, Fr_l = 4.5
85D, Fr_l = 4.5
0.2 0.2

0.1 0.1
r/b r/b
0 0
0 1 2 3 0 1 2 3
(a) Normalized RMS turbulent fluctuation (b) Normalized turbulent dissipation rate
/uc

Figure 6.5: CFD model predicted turbulent intensity and dissipation rate in the
jet and plume regime. F rl is the local densimetric Froude number defined as
Eq. 6.6. Solid symbols: plume regime; open symbols: jet regime.

147
6.2.2 The external flow
The entrainment by the buoyant jet acts as a sink to the surrounding fluid
and creates an external irrotational flow field. Previous studies (e.g. Sparks et
al. 1991) on vertical sediment-laden buoyant jet showed that, when the external
flow field is strong enough compared with the settling velocity of sediment par-
ticles, the vertical trajectory of the particle is deviated towards the jet, resulting
in re-entrainment and the change in deposition profile. In Chapter 5 the external
flow field is incorporated in the analytical solution of radial entrainment velocity
(Eqs. 5.6 and 6.13). However, in general buoyant jet situation, the entrainment
flow field is difficult to be determined analytically. Fig. 6.6a shows the external
flow field in a horizontal buoyant jet experiment predicted by the point sink ap-
proach, illustrating the complexity and three-dimensional nature of the external
flow. Fig. 6.6b show the particle velocity by superimposing the fluid velocity with
settling velocity. It clearly indicates that close to the jet, particles are attracted
back to the jet by the entrainment flow, while far away from the jet, the particle
velocities are less influenced and particles fall vertically.
The jet entrainment induced external flow is determined by the point sink
method introduced by Lai (2009) and Lai and Lee (2012a) (Fig. 6.7). The buoy-
ant jet is regarded as a number of point sinks (conveniently as each jet element
from the computation of JETLAG) along its trajectory. The strength of point
sink i at position (xi , yi , zi ) is equal to the entrainment flow per unit streamwise
length of the jet
dQi
mi = = 2bg g uc (6.15)
ds
where Qi is the jet volumetric flux
Qi = b2g uc . (6.16)
The velocity potential i induced by this jet element at an arbitrary location
(x, y, z) is given by
mi
i = s (6.17)
4r
q
where r = (x xi )2 + (y yi )2 + (z zi )2 is the distance from the sink.
The three-dimensional flow field (ui , vi , wi ) induced by this single jet element
can be given by differentiating the velocity potential with respect to x, y and z
directions.
i mi s r
ui = = (6.18)
x 4r2 x
mi (x xi )s
=
4 [(x xi ) + (y yi )2 + (z zi )2 ]3/2
2

i mi s r
vi = =
y 4r2 y
mi (y yi )s
=
4 [(x xi )2 + (y yi )2 + (z zi )2 ]3/2
i mi s r
wi = =
z 4r2 z

148
mi (z zi )s
=
4 [(x xi ) + (y yi )2 + (z zi )2 ]3/2
2

By summing up the induced velocities at a point (x, y, z) by all the sinks, the
total induced velocity by the jet entrainment can be found as (Lai, 2009)
N
X
u(x, y, z) = ui (x, y, z) (6.19)
i=1
XN
v(x, y, z) = vi (x, y, z)
i=1
XN
w(x, y, z) = wi (x, y, z)
i=1

where N is the number of jet elements. This represent the velocity field induced
by one jet. The induced flow determined by the above method is for an infinite
domain without any boundary. To account for the free surface and bottom
boundary, the method of images is used to compute the jet-induced flow, with
mirror images of the jet imposed for the bottom and surface boundaries.
It has to be noted that the potential flow theory is not valid inside the tur-
bulent jet, which is determined using the JETLAG solution. The location of the
transition between jet flow and external flow has to be determined. By compar-
ing the semi-analytical solution of the radial velocity using the point sink method
(assuming a solid boundary at the nozzle side) and the analytical radial velocity
from continuity equation Eq. 6.13 for a pure jet (Fig. 6.8), the velocity estimated
by point sink method is very close to that estimated using the analytical solu-
tion, for r > 3bg . It is adopted that the external flow is applied to the region
over three Gaussian jet widths from the jet centerline. Inside the jet (r < 3bg )
the velocity is calculated by Eqs. 6.11 and 6.13. Turbulence is negligible in the
external flow region.

149
0.4
0.01 m/s

0.2
z (m)

0 0.2 0.4 0.6 0.8 1


x (m)

(a) External flow predicted using point sink approach

0.4
0.02 m/s

0.2
z (m)

0 0.2 0.4 0.6 0.8 1


x (m)

(b) Sediment velocity for fine sand ws = 1.36cm/s, wp = wf ws

Figure 6.6: (a) The external flow field of a horizontal buoyant jet (u0 = 0.65m/s,
F r = 19.5) and (b) its influence on particle velocity. Note the different vector
scales

150
Figure 6.7: The point sink approach for determining the external flow (Lai,
2009).

0.02 0.02
x = 10D ur (m/s) x = 20D ur (m/s)
0.02 0.015
Point sink solution Point sink solution
0.01 Solution by continuity 0.01 Solution by continuity

0.01 0.005
r/b r/b
0.00 0
-4 -3 -2 -1 0 1 2 3 4 -4 -3 -2 -1 0 1 2 3 4
-0.01 -0.005

-0.01 -0.01

-0.02 -0.015

-0.02 -0.02

0.02
x = 35D ur (m/s)
0.015
Point sink solution
0.01 Solution by continuity

0.005
r/b
0
-4 -3 -2 -1 0 1 2 3 4
-0.005

-0.01

-0.015

-0.02

Figure 6.8: Comparison of the radial velocity using the point sink method (as-
suming a solid boundary for the plane x = 0) with the analytical solution
(Eq. 6.13).

151
6.2.3 The surface spreading layer
As the jet impinges the water surface, the excess pressure results in the jet fluid to
spread horizontally and radially in a layer. The dynamics of the spreading layer
is very complicated as it involves the transition from a near vertical turbulent
buoyant jet flow to a horizontal density-driven current. Internal hydraulic jump
may occur depending on the momentum and buoyancy flux of the jet. The
complex transitional flow field cannot be represented exactly. A simple integral
approach is adopted to calculate the spreading layer flow and the sedimentation.
This approach is based on the following assumptions.

The spreading current is steady.

The horizontal momentum induced by a non-vertical buoyant jet is negli-


gible and not included in the spreading layer calculation.

Integral model for spreading layer

Assuming azimuth symmetry on the radial velocity us (r) and buoyancy gs (r) =


g across the spreading layer with thickness hs (r) at a radial location of r from
the center of impingement (Fig. 6.9), the governing equations for the steady
spreading layer are:
Continuity:
dQs
= 2we r (6.20)
dr
Conservation of radial momentum:
!
dMs 1 d(rgs h2s )
= 2 i u2s r (6.21)
dr 2 dr
Conservation of buoyancy:
dBs
=0 (6.22)
dr
where

Qs = 2rhs us (6.23)
Ms = 2rhs u2s
Bs = 2rhs us gs

are the volumetric, kinematic momentum and specific buoyancy flux at radial
distance r from the center of impingement. we is the entrainment velocity from
the interface between the spreading current and the fluid below, given as a func-
tion of a Richardson number for the spreading layer (Pedersen, 1980; Akar and
Jirka, 1994).

we = 0.0015us Ri2 (6.24)


gs hs
Ri =
u2s

152
s (r0) s s s

Figure 6.9: The spreading current after jet impingement at the free surface.

153
i is the interfacial friction coefficient with values 0.002-0.005 (Abraham and
Eysink, 1971; Akar and Jirka, 1994) and taken as 0.003 in this study. The
variables hs , us and gs in governing equations can be solved numerically with
the initial conditions specified according to the prediction of JETLAG at the jet
terminal level.
The computation of JETLAG ceases as the upper top-hat boundary (xu , zu )
hits the water surface. The initial thickness hs (r0 ) of the spreading layer is
given by the vertical difference between the upper (zu ) and lower (zb ) top-hat
boundaries of the jet (hs (r0 ) = zu zb ) at the impingement. A minimum value
of hs (r0 ) = 0.08H is provided for the jet hitting the water surface near vertically,
according to experimental data and analytical prediction of vertical buoyant jets
(Lee and Jirka, 1981). The impingement point (xs , zs ) is defined as the jet
centerline position for which its jet upper boundary hits the water surface. The
initial flow rate Qs (r0 ) and buoyancy gs (r0 ) of the spreading current is essentially
taken as the jet volumetric flow rate and buoyancy at the terminal level. The
initial radius r0 is taken as the horizontal component of the top hat width of the
jet at the impingement. The initial radial velocity us (r0 ) can thus be estimated.
Qs (r0 )
us (r0 ) = (6.25)
2r0 hs (r0 )
Across the spreading layer, the vertical velocity distribution is assumed as
half-Gaussian with the surface at its maximum,
!
z2
us (r, z) = usm exp 2 (6.26)
hsm
where z is the distance from the free surface. By continuity and conservation of
momentum, the maximum velocity usm and Gaussian thickness hsm (r) can be
related to the average values us and hs by

usm = 2us (6.27)
s
2
hsm = hs (6.28)

For r < r0 (the transition zone), a quadratic relation is used to interpolate
the radial velocity from zero at the impingement to the maximum value at the
jet boundary r0 . A Gaussian profile is adopted for the vertical velocity, reducing
from the jet velocity to zero at the surface quadratically.
A CFD model is developed to predict the surface spreading current induced
by a vertical buoyant jet and validate the integral model. Details of the model
is given in Appendix B. Two buoyant jet conditions are simulated: Case A:
u0 = 0.351m/s, D = 7mm and F r = 10.0; Case B: u0 = 0.146m/s, D = 7mm
and F r = 4.0. Height of the jet H = 0.3m. In the integral model, the initial
condition at r = r0 is determined from the corresponding JETLAG simulation of
the buoyant jet. Fig. 6.10 shows the comparison between CFD and integral
models for the mean spreading layer current, thickness and reduced gravity.
The integral model successfully captures the variation of the spreading layer
dynamics.

154
0.06 0.06
ur (m/s) hs (m)
0.05 0.05 Integral Model
Integral Model
0.04 CFD 0.04

0.03 0.03

0.02 0.02

0.01 0.01
r (m) r (m)
0 0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4

(i) Velocity (ii) Thickness


0.02
2
g' (m/s )
Integral Model
CFD

0.01

r (m)
0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4

(iii) Reduced gravity


(a) Case A: u0 = 0.351m/s, D = 7mm, F r = 10.0

0.06 0.06
usm (m/s) hs (m)
0.05 0.05 Integral Model
Integral Model
0.04 CFD 0.04

0.03 0.03

0.02 0.02

0.01 0.01
r (m) r (m)
0 0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4

(i) Velocity (ii) Thickness


0.02
2
g' (m/s )
Integral Model
CFD

0.01

0
r (m)
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4

(iii) Reduced gravity


(b) Case B: u0 = 0.146m/s, D = 7mm, F r = 4.0

Figure 6.10: Comparison of integral model and CFD predicted mean spreading
current velocity us , thickness hs and mean reduced gravity gs .

155
6.3 Particle tracking model
The flow field of an arbitrarily inclined buoyant jet in stagnant water can be
determined readily using the semi-analytical methods described in the previous
section. Any sediment particles introduced to the flow field can hence be tracked
for its position without difficulty using the particle tracking model developed in
Chapter 3. The simplified model, assuming particle velocity as the sum of fluid
and stillwater settling velocity, is extensively used for all cases reported hereafter.
The full model, solving the governing equation of particle motion, is utilized for
sensitivity analysis for experimental cases using particles other than sand/glass
(Neves and Fernando, 1995; Ernst et al. 1996). The importance of the Basset
force is evaluated for these independent studies on vertical sediment-laden jets
(Section 6.4.4).
For each case, the mean flow field is pre-computed. Jet turbulent velocity
fluctuation and turbulence length and time scales are determined from best-
fitted self-similar profiles of RMS turbulent velocity and turbulent kinetic energy
dissipation rate from CFD solutions of single-phase vertical buoyant jets. Par-
ticles are introduced at the end of the zone of flow establishment and tracked
until they reached the tank bottom level. In these cases, Np = 10, 000 particles
are used to obtain the deposition profile.

156
6.4 Vertically upward sediment-laden buoyant
jet

6.4.1 General observation


The vertically upward sediment-laden buoyant jet is an important special case
for studying the sediment deposition from general buoyant jets. Extensive ex-
perimental studies and theoretical modelling have been documented (e.g. Carey
et al. 1988; Sparks et al. 1991; Neves and Fernando, 1995; Ernst et al. 1996;
Zarrebini and Cardoso, 2000; Cardoso and Zarrebini, 2002). Most of these stud-
ies aim at explaining and predicting the ash deposition from volcanic eruption.
Sediment particles are mixed with jet fluid and ejected by the jet flow upwards.
When the jet velocity is sufficient high (wj >> ws ), the sediment particles are
carried upwards. Jet velocity decays due to turbulent momentum transfer with
ambient fluid. Two regimes of particle fall-out is in general observed in these
past studies:

1. As jet velocity reduces to a level similar to the settling velocity of particles


(wj ws ), the sediment is no longer carried upwards but reaches its max-
imum height of rise. Particles near the edge of the plume, will first settle
out along the jet boundary while those near the centerline is pushed out
to the edge. Particles fall to the bottom near the plume edge with slightly
curved trajectories. Some particles are re-entrained to the plume and some
are recycled several times before they deposit (Neves and Fernando, 1995;
Ernst et al. 1996). (Fig. 6.11a)

2. If the fluid velocity is sufficiently high to carry the particle until it reaches
the water surface, particles are transported laterally to the horizontal
spreading current. In the spreading current, turbulence maintains the par-
ticles in suspension, until it is not strong enough to carry the particles and
they fall out from the interface of the spreading current and the ambient
water. The falling particles follow a curved trajectory towards the nozzle
due to the ambient entrainment flow. Re-entrainment occurs when the
particle fall out near the edge. A region of re-entrainment close to the
plume can be observed where particles must be re-entrained as they fall
out from the spreading current. The size of the region is characterised by
a critical radius which depends on the strength of the gravity current and
the settling velocity of particles (Sparks et al. 1991; Zarrebini and Cardoso,
2000; Cardoso and Zarrebini, 2002). (Fig. 6.11b)

Particle model simulations are carried out to compare with experimental data
of these studies. The numerical results of Section 6.4.2 and 6.4.3 are obtained
using the Simplified Model. Since the particle properties in these independent
experiments are different from those used in the present study, the Full Model is
also utilized to investigate the validity of the simplified model. The comparison
of model results using both models are presented in Section 6.4.4.

157
zm

(a) Fall out from jet margin

(b) Fall out from spreading current

Figure 6.11: The two fall out mechanisms from vertical sediment-laden buoyant
jet.

158
6.4.2 Sediment fall out from jet margin
Neves and Fernando (1995)

The experiments were carried out with pure momentum jets, in a 0.4 0.4
0.6m deep tank with a jet nozzle diameter of 5.6mm. Sediment are light
polystyrene particles of diameter d = 530m (p = 1025.1kg/m3 ), 799m (p =
1044.5kg/m3 ) and 868m (p = 1044.5kg/m3 ). Settling velocity ws is estimated
to be 0.33, 0.97 and 1.09 cm/s respectively. Initial concentration of sediments
ranges from 0.0045%-0.23% by volume fraction, which are very dilute conditions
similar to those used in the present study. Sediment deposition is collected using
annular trays circumfencing the jet nozzle and measured by manual counting of
particles.
Details of experimental jet flow conditions were not provided but expres-
sions for estimating the maximum height of rise, deposition profile, maximum
deposition rate and its location were presented based on best-fitting of their
experimental data. The expressions are expressed in terms of the momentum-
1/2
settling length scale lm = M0 /ws and a particle inertia-buoyancy length scale
lb = ws d2 / = Rep d. lb is the length scale for which a particle to reach its ter-
minal velocity, which is very short compared with lm and the distance travelled
by the particle in the jet (see Table 6.1). Their results shows that:

Deposition profile:
! !2
F r r
= 65 exp 0.054 + 7.8 (6.29)
Fmax rmax
rmax

r = r 0.5D

Maximum rise of particles:


zm = 6.9lm (6.30)

Maximum deposition rate (mass per unit area)


2
Fmax = 1.4Q0 C0 /lm (6.31)

Location of maximum deposition rate



rmax = rmax 0.5D = 0.15lm (6.32)

To justify the particle tracking model in predicting the deposition of upward


jets, several numerical test experiments are designed based on the set-up of Neves
and Fernando (1995). Constrained by the tank depth of 0.5m and the maxi-
mum height of particle rise given by Eq. 6.30, the jet velocity in the numerical
experiments cannot exceed 0.15 m/s for particles of 799m and 868m , and
0.05 m/s for particles of 530m . Table 6.1 shows the numerical experiment
conditions. It can be seen that lm is 1 order greater than lb . The deposition

159
Table 6.1: Numerical experiments for vertical sediment-laden pure jets (Neves
and Fernando, 1995).
Case Jet Sediment Sediment Settling lm = lb = z(ws = wc )
1/2
velocity density diameter velocitya M0 /ws ws d2 /
m/s g/cm3 m cm/s cm cm cm
NF530-1 0.030 1.025 530 0.33 4.51 0.093 23.9
NF530-2 0.040 1.025 530 0.33 6.02 0.093 31.9
NF530-3 0.050 1.025 530 0.33 7.52 0.093 47.8
NF799-1 0.075 1.045 799 0.97 3.84 0.619 26.8
NF799-2 0.100 1.045 799 0.97 5.12 0.619 35.8
NF799-3 0.150 1.045 799 0.97 7.67 0.619 53.7
NF868-1 0.075 1.045 868 1.09 3.41 0.821 31.6
NF868-2 0.100 1.045 868 1.09 4.55 0.821 42.1
NF868-3 0.150 1.045 868 1.09 6.83 0.821 52.6

Jet diameter = 5.6mm


Sediment concentration = 0.05g/L
wc = jet centerline velocity
a
Particle settling velocity in 20 C water, estimated based on spherical particle drag law.

profiles are obtained from summing the mass of particles in concentric regions of
gap width 5mm and divided by the area of the region.
The comparison between model predicted and experiment measured deposi-
tion rate around the jet nozzle is very good for all the cases, showing a profile of
a peak deposition rate close to the jet nozzle and a long tail behind it (Fig. 6.12).
The numerical predictions collapse into a similar curve after normalizing against
the maximum values according to Eq. 6.29. The maximum height of rise of each
experiment zm can be treated as a visual upper boundary of the sediment-laden
jet, which is determined by the mean zm of all particles plus 2 standard devia-
tions. It is about equal to the level where ws = wc , where wc is the centerline
jet velocity. As some particles are laterally transported by turbulent fluctuations
and the fluid velocity away from the centerline is not sufficient to support further
upward transport of sediments, not all particles travel to zm . Fig. 6.13 shows
the linear relationship between zm and lm . The fitted constant of 6.77 is very
close to the value of 6.9 in Eq. 6.30, which indicates that the model predicts
very well the dynamics of the particles in turbulent jet flow. Fig. 6.14 shows
the normalized maximum deposition rate and its radial location by the model
prediction. The maximum deposition is about twice that than of Eq. 6.31 (3 vs
1.4). The location of maximum deposition is similar (Eq. 6.32, 0.1 vs 0.15). It
has to be noted that the deposition measurement close to the jet nozzle is highly
uncertain. Overall the comparison is satisfactory.

Ernst et al. (1996)

The experiments were carried out in a tank of 1.2m1.2m0.5m depth with


a jet nozzle diameter of 4mm. Freshwater jets with various initial velocity are
directed to the tank filled with salt water with various density to produce different

160
1.2 F/Fmax

530m u0 = 0.03m/s
1 u0 = 0.04m/s
u0 = 0.05m/s
0.8
N&F (1995)

0.6

0.4

0.2

0
0 1 2 3 4 5 6 7 8
r'/r max'
(a) 530m
1.2
F/Fmax u0 = 0.075m/s
1
799m u0 = 0.1m/s
u0 = 0.15m/s
0.8 N&F (1995)

0.6

0.4

0.2
r'/r max'
0
0 1 2 3 4 5 6 7 8

(b) 799m
1.2
F/Fmax u0 = 0.075m/s
1 868m u0 = 0.10m/s
u0 = 0.15m/s
0.8 N&F (1995)

0.6

0.4

0.2
r'/rmax'
0
0 1 2 3 4 5 6 7 8

(c) 868m

Figure 6.12: Comparison of model prediction of bottom deposition of vertical


sediment-laden jets and experimental data of Neves and Fernando (1995).

161
zm (m)
0.6

0.5

0.4 y = 6.7704x

0.3

0.2

0.1

0
0 0.05 l m (m) 0.1

Figure 6.13: Relation between lm and the maximum height of rise zm , model
prediction for vertical sediment-laden jet experiments of Neves and Fernando
(1995).

Fmax/(Q0C0/lm 2)
10

Model

1 N&F
530um
799um
868um
0.1
0.01 0.1 lb/lm 1
(a) Maximum deposition rate Fmax

1
rmax'/lm

N&F
0.1
Model 530um
799um
868um
0.01
0.01 0.1 lb/lm 1

(b) Location of maximum deposition rate rmax

Figure 6.14: Relation of the maximum deposition rate Fmax and position rmax

with lm in vertical sediment-laden jet experiments. The constant of 3.0 and 0.1
are compared with the value of 1.4 and 0.15 given by Neves and Fernando (1995)
(Eqs. 6.31 and 6.32.

162
buoyancy conditions. Sediment used is silicon carbide particles with a density
of 3210 kg/m3 . Sediment concentration ranges from <0.2 g/L to 2.5 g/L (<
0.08% by volume) to avoid influence of sediment-induced buoyancy. A total of
17 experiments are reported in the paper in which the results of 10 of them are
presented. Eight experiments are used for the present study and their parameters
are shown in Table 6.2. The first four are jet-like cases with high F r while the
other four are plume-like cases. All bottom deposition data presented in the
paper is normalized with the maximum accumulated mass deposition near the
nozzle.

Table 6.2: Vertical buoyant jet experiments (fall out from jet margin) in Ernst
et al. (1996)
3/4
M0
Case Jet initial Jet sediment ls = 1/2 Fr Sediment Settling
B0
vel. (m/s) conc. (g/L) (cm) = uj dia. (m ) vel.a (cm/s)
g0 D
Jet cases:
406 0.4775 0.2 8.1 21.4 275.0 4.3
407 0.7479 0.4 12.9 34.2 275.0 4.3
709 0.7479 <0.4 12.3 32.6 134.5 2.1
710 0.7479 <<0.2 12.2 32.4 327.5 5.3
Plume cases:
327 0.1512 <1 2.49 6.6 193.5 2.9
402 0.0796 <<2.4 1.33 3.5 193.5 2.9
404 0.0796 <<1.7 1.44 3.8 115.0 1.6
805 0.1512 <<0.8 1.15 3.0 193.5 2.9

Jet diameter = 4mm


a
Settling velocity as in Ernst et al. (1996).

Comparisons of model prediction and experimental data are shown in Fig. 6.15
for the four jet-like cases of Ernst et al. (1996). The comparison is excellent for
all the cases, showing a peak deposition close to the jet nozzle consistent with
the observation. The higher the jet flow, the larger the extent of the deposition
profile is (Case 406 vs 407). With the same jet flow, the extent of the deposi-
tion profile is smaller for larger settling velocity (Cases 709 vs 710). The lighter
particle and/or stronger jet flow result in the expelling of particles further away
from the jet and particles deposit in a wider region.
Results for the plume-like cases are shown in Fig. 6.16. Compared with the
jet like cases, it is observed that the maximum deposition is much closer to
the jet nozzle and the bottom deposition profile decays much faster, because
of the higher entrainment velocity induced by plumes (entrainment coefficient
g = 0.057 for a pure jet vs 0.088 for a pure plume). This is consistent with
the conclusion of Ernst et al. (1996) that the decay of deposition rate in radial
direction follows r1 for jet and r1/3 for plume.

163
1.2 1.2
F/Fmax F/Fmax
Prediction 407 Prediction
1 406 1
Measurement Measurement
0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2
r (m) r (m)
0 0
0 0.01 0.02 0.03 0.04 0.05 0 0.01 0.02 0.03 0.04 0.05

(a) Case 406: u0 = 0.48m/s, Fr = 21.4, ws = (b) Case 407: u0 = 0.75m/s, Fr = 34.2, ws =
4.3 cm/s 4.3 cm/s
1.2 1.2
F/Fmax F/Fmax
709 Prediction 710 Prediction
1 1
Measurement
0.8 0.8 Measurement

0.6 0.6

0.4 0.4

0.2 0.2
r (m) r (m)
0 0
0 0.01 0.02 0.03 0.04 0.05 0 0.01 0.02 0.03 0.04 0.05

(c) Case 709: u0 = 0.75m/s, Fr = 32.6, ws = (d) Case 710: u0 = 0.75m/s, Fr = 32.4, ws =
2.1 cm/s 5.3 cm/s

Figure 6.15: Comparison of predicted and measured radial sediment deposition


profile of the four vertical sediment-laden jet cases of Ernst et al. (1996). r is the
radial distance from the nozzle. F/Fmax is the deposition (g/m2 /s) normalized
against the maximum predicted/measured values.

1.2 1.2
F/Fmax F/Fmax
327 402
1 Prediction 1 Prediction

0.8 Measurement 0.8 Measurement

0.6 0.6

0.4 0.4

0.2 0.2
r (m) r (m)
0 0
0 0.01 0.02 0.03 0.04 0.05 0 0.01 0.02 0.03 0.04 0.05

(a) Case 327: u0 = 0.15m/s, Fr = 6.6, ws = 2.9 (b) Case 402: u0 = 0.08m/s, Fr = 3.5, ws = 2.9
cm/s cm/s
1.2 1.2
F/Fmax F/Fmax
404 805
1 Prediction 1 Prediction

0.8 Measurement 0.8 Measurement

0.6 0.6

0.4 0.4

0.2 0.2
r (m) r (m)
0 0
0 0.01 0.02 0.03 0.04 0.05 0 0.01 0.02 0.03 0.04 0.05

(c) Case 404: u0 = 0.08m/s, Fr = 3.8, ws = 1.6 (d) Case 805: u0 = 0.15m/s, Fr = 3.0, ws = 2.9
cm/s cm/s

Figure 6.16: Comparison of predicted and measured radial sediment deposition


profile of the four vertical sediment-laden plume cases of Ernst et al. (1996).
r is the radial distance from the nozzle. F/Fmax is the deposition (g/m2 /s)
normalized against the maximum predicted/measured values.

164
6.4.3 Sediment fall out from spreading current
Sparks et al. (1991)

The experiments were carried out in a tank of 1.2m1.2m0.46m depth with


a jet nozzle diameter of 8mm. Freshwater jets with u0 = 0.0215m/s (Q0 =
1.08cm3 /s) are directed to salt water with density 1021 kg/m3 in the tank. The
densimetric Froude number is about 0.5, representing a pure plume. The present
model independently computes the potential core development and accounts for
the contraction and velocity acceleration due to buoyancy (Lee and Chu, 2003).
Non-spherical silicon carbide particles with a density of 3210 kg/m3 are used
as sediment seedings with concentration 10 g/L (0.3% by volume). Sediment
diameter d ranges from 28m to 131m . Settling velocity ws is estimated using
Hallermeier (1981)s formula, same as that in Sparks et al. (1991):

ws = D3 /18d for D3 < 39 (6.33)


ws = D2.1 /6d for 39 < D3 < 104
ws = 1.05D1.5 /d for 104 < D3 < 3 106
h i1/3
D = g(s 1)/ 2 d

The characteristics of three experiments used in this study are shown in Table
6.3.

Zarrebini and Cardoso (2000)

The experiments were carried out in a tank of 0.75m0.75m0.3m depth with a


jet nozzle diameter of 7mm. Freshwater jets with various flow rates are directed
to salt water in the tank to produce different buoyancy flux. The densimetric
Froude number ranges from 4.0 to 10.5. Spherical Ballotini particles with a
density of 2470 kg/m3 are used with sediment concentration 5-6 g/L ( 0.2% by
volume). Sediment sizes ranges from 49m to 81m . The characteristics of the
one experiment used in the present study is shown in Table 6.3.

Results

The comparison between predicted and measured deposition profiles are shown
in Fig. 6.17. The deposition is a cone-like shape with a maximum close to the
jet nozzle. Compared with the fall out from jet margin, the sediment fall out
from the spreading current produces a much larger region of deposition, extent to
over 0.1-0.4m, depending on the particle size and the height of jet impingement.
Sparks et al. (1991) derived that the decay of the deposition follows the r2 law,
representing a much larger deposition extent than the r1 law for jet and r1/3
law for plume. The model predicted deposition agrees well with a wide range of
particle size (57-96m ). The reason for the discrepancy in Sparks et al. (1991)s
experiment using 67m particles (Fig. 6.17b) is not clear.

165
Table 6.3: Vertical buoyant jet experiments (fall out from surface current) of
Sparks et al. (1991), and Zarrebini and Cardoso (2000)
Case Jet initial Jet sediment Fr Sediment Settling
u
vel. (m/s) conc. (g/L) = j dia. (m ) velocity (mm/s)
g0 D
Sparks et al. (1991)
83-2 0.0215 10 0.53 58 4.03a
810-2 0.0215 10 0.53 67 5.38a
86-1 0.0215 10 0.53 96 11.05a
Zarrebini and Cardoso (2000)
13 0.3513 6 10.0 65.03 3.34b

Jet diameter - Sparks et al. (1991): 8mm; Zarrebini and Cardoso (2000): 7mm
a
Predicted settling velocity using Hallermeier (1981)s formula (Eq. 6.33), in 20 C water.
b
Predicted settling velocity using spherical drag law, in 20 C water.

F/Fmax F/Fmax
1.2 1.2
Model Model
1 Expt. (Sparks) 1 Expt. (Sparks)
57m 67m
0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2
r (m) r (m)
0 0
0 0.05 0.1 0.15 0.2 0.25 0.3 0 0.05 0.1 0.15 0.2 0.25

(a) Sparks et al. (1991) Expt 83-2, u0 = (b) Sparks et al. (1991) Expt 810-2, u0 =
0.0215m/s, F r = 0.53, d = 57m 0.0215m/s, F r = 0.53, d = 67m

F/Fmax F/Fmax
1.2 1.2
Model Model
1 Expt. (Sparks) 1 Expt 13. (ZC2000)
96m 65m
0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2
r (m) r (m)
0 0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0 0.1 0.2 0.3 0.4 0.5

(c) Sparks et al. (1991) Expt 86-1, u0 = (d) Zarrebini and Cardoso (2000) Expt 13,
0.0215m/s, F r = 0.53, d = 96m u0 = 0.35m/s, F r = 10, d = 65m

Figure 6.17: Comparison of numerical model prediction and experimental data of


vertical sediment jet experiments: (a)-(c) Sparks et al. (1991), (d) Zarrebini and
Cardoso (2000). r is the radial distance from the nozzle. F/Fmax is the deposition
(g/m2 /s) normalized against the maximum predicted/measured values.

166
6.4.4 Comparison of the Full Model and the Simplified
Model
Particles in the independent investigations of vertical upward sediment-laden jets
are different from those used in the present study: Neves and Fernando (1995)
used polystyrene particles of p = 1040 kg/m3 ; Ernst et al. (1996) and Sparks et
al. (1991) used silicon carbide particles of 3210 kg/m3 . The full model is used to
simulate selected cases to justify the importance of the Basset force and the use
of simplified model on these particles.
Fig. 6.18 shows the comparison of Full Model (both with and without the
Basset force) and the Simplified Model on predicting the bottom deposition of
vertical jet cases of Neves and Fernando (1995) and Ernst et al. (1996). Surpris-
ingly, the both model works similarly well for vertical jets for the light polystyrene
particles used by Neves and Fernando (1995), contrary to the previous hori-
zontal jet experiments which the Full Model is required for plastic particles of
specific gravity of 1.1 and 1.5. The initial jet velocity for these experiments
(u0 = 0.03 0.15m/s, Re = 168 840) are smaller than the present horizontal
jet experiments (u0 = 0.3 0.9m/s). The turbulent intensity is smaller such that
even these light particle can follow the fluid motion. Both models also works
well for the high density silicon carbide particles used by Ernst et al. (1996).
Similar to the previous sensitivity analysis (Chapter 5, Section 5.6.3), Basset
force has little importance in the prediction of deposition profiles, thus can be
tacitly ignored.

167
2 2
0.05 F (g/m /s) 0.16 F (g/m /s)
Full model w/o Basset Full model w/o Basset
0.14
0.04 Full model w/ Basset Full model w/ Basset
0.12
Simp. Model Simp. Model
0.03 0.1
530 m 868 m
0.08
0.02 0.06
0.04
0.01
r (m) 0.02
r (m)
0 0
0 0.01 0.02 0.03 0.04 0.05 0 0.01 0.02 0.03 0.04 0.05

3 3
(a) NF530-3: u0 = 0.05 m/s, s = 1025 kg/m , (b) NF868-3: u0 = 0.15 m/s, s = 1045 kg/m ,
d = 530m, ws = 1.09 cm/s d = 868m, ws = 0.33 cm/s

1.2 1.2
F/Fmax Full Model w/o Basset F/Fmax Full Model w/o Basset
710 406
1 Full Model w/ Basset 1 Full Model w/ Basset
Simp. Model Simp. Model
0.8 0.8 Meas.
Meas.
0.6 0.6

0.4 0.4

0.2 0.2
r (m) r (m)
0 0
0 0.01 0.02 0.03 0.04 0.05 0 0.01 0.02 0.03 0.04 0.05

(c) Case 710: u0 = 0.75m/s, Fr = 32.4, s = (d) Case 406: u0 = 0.48m/s, Fr = 21.4, s =
3210 kg/m3, d = 328m, ws = 5.3 cm/s 3210 kg/m3, d = 275m, ws = 4.3 cm/s

1.2 1.2
F/Fmax F/Fmax
402 Full Model w/o Basset 404 Full Model w/o Basset
1 1
Full Model w/ Basset Full Model w/ Basset
0.8 Simp. Model 0.8 Simp. Model
Meas. Meas.
0.6 0.6

0.4 0.4

0.2 0.2
r (m) r (m)
0 0
0 0.01 0.02 0.03 0.04 0.05 0 0.01 0.02 0.03 0.04 0.05

(e) Case 402: u0 = 0.08m/s, Fr = 3.5, s = (f) Case 404: u0 = 0.08m/s, Fr = 3.8, s =
3210 kg/m3, d = 194m, ws = 2.9 cm/s 3210 kg/m3, d = 115m, ws = 1.6 cm/s

Figure 6.18: Comparison of the Full Model (with and without Basset force) and
Simplified Model for vertical sediment jet experiments of (a)-(b): Neves and
Fernando (1995) (p = 1040 kg/m3 ); (c)-(f): Ernst et al. (1996) (p = 3210
kg/m3 ).

168
6.5 Horizontal sediment-laden buoyant jet
6.5.1 Experiments of Li (2006) and Lee (2010)
Li (2006) and Lee (2010) investigated the dynamics of sediment-laden horizontal
buoyant jet and its bottom deposition profile with extensive experiments. The
experimental data cover a wide range of buoyant jet conditions and sediment
characteristics (Tables 6.4, 6.5 and 6.6). They provide a comprehensive dataset
for validating the present particle tracking model.
The experiments were carried out in a tank similar to the present experiments,
in the size of 1m1m0.45m depth and filled with freshwater. The jet nozzle
was 0.04m above the top of the sediment collecting tray and 0.38m below the
water surface. The jet fluid was a mixture of freshwater and ethanol in different
volume fractions to simulate jets with different buoyancy flux. Sediment was
introduced to the buoyant jet flow through an hourglass to provide constant
sediment concentration to the jet. Sediment concentration ranges from 0.8 g/L
to 6.9 g/L (0.03%-0.28% by volume). Sediment in such dilute concentration
does not affect the fluid-phase properties significantly. Natural sand particles
(p = 2.65 g/cm3 , Coarse Sand: deq = 166m , Fine Sand: deq = 133m ) are
used in experiments of Li (2006). Spherical glass particles (p = 2.50 g/cm3 ,
S199, S153: d50 = 199, 153m , G215, G180, G115: d50 = 215, 180, 115m )
were used in experiments of Lee (2010). The longitudinal deposition profiles (in
g/m/s) were measured.
In Tables 6.4, 6.5 and 6.6, lm refers to the momentum-settling length scale
defined as Eq. 5.1 and ls is the momentum-buoyancy length scale (Eq. 6.2). For
a horizontal buoyant jet, its trajectory is significantly curved upwards by the
buoyancy beyond x 3ls (Fischer et al. 1979; Lee and Chu, 2003). There are
two modes of bottom deposition in horizontal sediment-laden buoyant jets - the
fall out from jet region and fall out from spreading current. x lm represents
the location of the maximum deposition of fall out from jet region, while and
x 3ls represents the the location of the maximum deposition of fall out from
spreading current. The two length scales compare the importance of the two
sedimentation regions. When lm < 3ls , most sediment falls out from the jet
region before it is significantly curved upwards, similar to a pure momentum
horizontal jet. Little sediment is transported to the surface layer, resulting in
a significant first peak and a very small second peak. When lm > 3ls , a large
proportion of sediment is transported to the surface spreading layer before falling
out from the jet, forming a double peak deposition profile. The significance of
these length scales is discussed in the model comparison section.

169
Table 6.4: Horizontal buoyant jet experiments of Li (2006).
Case u0 C0 ws0 wsa a 0 Fr lm 3ls
(m/s) (g/L) (cm/s) (cm/s) (kg/m3 ) (kg/m3 ) (m) (m)
Coarse sand, deq = 166m
CB66 0.6484 3.2358 1.298 1.947 998.2 979.4 19.47 0.2656 0.3300
CB62 0.6091 3.2700 1.332 1.947 998.2 980.4 18.80 0.2432 0.3186
CB54 0.5305 3.6347 1.648 1.947 998.2 980.4 16.37 0.1712 0.2775
CB50 0.4912 4.2000 1.308 1.947 998.2 979.7 14.87 0.1997 0.2520
Fine sand, deq = 133m
FB66 0.6484 3.5960 0.940 1.358 998.0 981.5 20.79 0.3668 0.3522
FB58 0.5698 4.0780 0.741 1.358 998.0 971.8 14.50 0.4089 0.2456
FB46 0.4519 4.9000 0.940 1.358 998.0 981.5 14.49 0.2556 0.2455

Jet diameter = 6mm


1/2
lm = M0 /ws = (/4)1/2 u0 D/ws
3/4 1/2
ls = M0 /B0 = (/4)1/4 F rD

Table 6.5: Horizontal buoyant jet experiments of Lee (2010), S199 and S153
particles.
Case u0 C0 ws0 wsa a 0 Fr lm 3ls
(m/s) (g/L) (cm/s) (cm/s) (kg/m3 ) (kg/m3 ) (m) (m)
S199 particles, d50 = 199m
S199B46-Fr9 0.4519 5.73 1.24 2.20 998.3 962.6 9.85 0.1092 0.1669
S199B58-Fr12 0.5698 3.31 1.32 2.24 998.0 960.9 12.18 0.1353 0.2064
S199B50-Fr14 0.4912 4.46 1.60 2.21 997.4 978.3 14.63 0.1182 0.2479
S153 particles, d50 = 153m
S153B46-Fr9 0.4519 4.06 0.77 1.50 998.0 961.8 9.78 0.1602 0.1657
S153B52-Fr11 0.5109 1.94 0.77 1.48 997.7 962.4 11.19 0.1835 0.1897
S153B58-Fr12 0.5698 2.72 0.76 1.50 998.0 960.7 12.15 0.2020 0.2059

Jet diameter = 6mm


1/2
lm = M0 /ws = (/4)1/2 u0 D/ws
3/4 1/2
ls = M0 /B0 = (/4)1/4 F rD

170
Table 6.6: Horizontal buoyant jet experiments of Lee (2010), G215, G180 and
G115 particles.
Case u0 C0 ws0 wsa a 0 Fr lm 3ls
(m/s) (g/L) (cm/s) (cm/s) (kg/m3 ) (kg/m3 ) (m) (m)
G215 particles, d50 = 215m
G215B50-Fr15 0.4912 6.90 2.07 2.66 997.1 980.1 15.51 0.0982 0.2628
G215B50-Fr12 0.4912 6.71 1.82 2.69 996.3 969.9 12.44 0.0971 0.2108
G215B50-Fr10 0.4912 6.73 1.56 2.68 996.7 956.5 10.08 0.0975 0.1708
G215B90-Fr27 0.8842 3.84 2.09 2.68 997.0 980.0 27.91 0.1754 0.4729
G215B90-Fr22 0.8842 3.78 1.81 2.68 996.4 970.3 22.52 0.1754 0.3816
G215B90-Fr17 0.8842 3.81 1.56 2.67 996.3 956.3 18.19 0.1761 0.3082
G180 particles, d50 = 180m
G180B50-Fr15 0.4912 3.80 1.57 2.06 997.1 980.0 15.46 0.1268 0.2620
G180B50-Fr12 0.4912 3.92 1.34 2.06 996.8 970.4 12.44 0.1268 0.2108
G180B50-Fr10 0.4912 3.84 1.16 2.07 996.3 955.8 10.04 0.1262 0.1702
G180B90-Fr27 0.8842 2.07 1.58 2.06 997.1 980.5 28.25 0.2282 0.4786
G180B90-Fr22 0.8842 2.16 1.33 2.05 996.3 970.5 22.65 0.2293 0.3838
G180B90-Fr17 0.8842 2.12 1.16 2.06 996.1 956.6 18.30 0.2282 0.3101
G115 particles, d50 = 115m
G115B50-Fr15 0.4912 1.34 0.75 1.02 996.6 980.3 15.83 0.2561 0.2683
G115B50-Fr12 0.4912 1.35 0.61 1.01 996.5 970.4 12.51 0.2586 0.2120
G115B50-Fr10 0.4912 1.34 0.55 1.00 996.2 957.1 10.22 0.2612 0.1732
G115B90-Fr28 0.8842 0.79 0.73 1.00 996.6 980.1 28.32 0.4702 0.4800
G115B90-Fr22 0.8842 0.77 0.61 1.00 996.1 970.5 22.73 0.4702 0.3852
G115B90-Fr17 0.8842 0.75 0.54 1.00 996.3 956.6 18.26 0.4702 0.3094

Jet diameter = 6mm


1/2
lm = M0 /ws = (/4)1/2 u0 D/ws
3/4 1/2
ls = M0 /B0 = (/4)1/4 F rD

171
6.5.2 CFD model prediction

CFD model is developed to study the bottom deposition of horizontal sediment-


laden buoyant jets. Details of model and the comparison on single phase jets is
discussed in Appendix B. The model grid and boundary conditions are similar
to the one used for horizontal momentum jet. It should be noted that the water
surface is represented by a symmetric boundary to account for the zero gradient
condition there. Except for the bottom and the wall which the jet nozzle is at,
the other three vertical walls are set as open boundaries to allow for a steady
state solution. Since the particle settling velocities in ethanol-water mixture
and in freshwater are significantly different (see Tables 6.4-6.6), instead of us-
ing a single settling velocity, the local settling velocity, evaluated using Soulsby
(1997)s equation (Eq. 5.10) based on the local computed density and molecu-
lar viscosity is used for calculating the settling flux in the sediment transport
equation in these buoyant jet cases. The temperature-dependent density and
molecular viscosity of water-ethanol mixtures is shown in Appendix C. In the
model, the local fluid density is predicted, the percentage of ethanol by weight is
then back-estimated to obtain the molecular viscosity of the fluid. An empirical
correction factor of 0.75 is applied to account for the settling velocity reduction
by turbulence, similar to that in momentum jets. A total of 7 experiments of
particle-laden horizontal buoyant jets from Li (2006), involving fine and coarse
sand particles, are studied using CFD (Table 6.4).
The CFD predicted sediment deposition profiles are compared with the ex-
perimental results (Fig. 6.19). The comparison is very good for coarse sands
(Fig. 6.19a), showing clearly the two deposition pattern: the sharp peak caused
by fall out from the momentum dominated region near the jet exit and a slowly
decaying deposition region from the gravitational current. The first peak is the
dominant deposition model for coarse particle jets. For the cases of fine sand
(Fig. 6.19b), the computed profile is less comparable to the measured deposition
profiles, nevertheless the general two-peak characteristics of the profiles can still
be predicted. As the sediment are mainly fall out from the gravitational spread-
ing current, the deposition is significantly affected by the large region outside the
jet which the flow is induced by the jet entrainment. The deposition prediction
is prone to model uncertainties in the CFD prediction of the external current,
spreading current and turbulence intensity.
Figs. 6.20 and 6.21 show the predicted sediment concentration profiles in the
centerline plane, with the conservative tracer concentration profiles side by side.
Due to the effect of settling, the sediment concentration profiles migrates towards
the lower edge of the jet. The maximum sediment concentration is always located
in the lower half of the buoyant jet, unlike the tracer concentration maximum
following the jet centerline. The coarse particles show significant drop out prior
to the curve up of the buoyant jet, leading to very little sediment rising up with
the buoyant jet to the top spreading current. For fine sand, the particles could
rise up to the spreading current but through the lower half of the jet.
To summarise, CFD model predicts the bottom deposition of sediment-laden
buoyant jets reasonably, provided that the settling velocity is empirically cor-

172
rected according to the turbulent level. Nevertheless, the prediction is subjected
to the model inaccuracy in external flow and turbulent level prediction.
0.4 0.4
CB66 CB62
FLUENT
FLUENT
0.3 0.3 Measured
Measured
Fs (g/m/s)

Fs (g/m/s)
0.2 0.2

0.1 0.1

x (m) x (m)
0 0
0 0.2 0.4 0.6 0 0.2 0.4 0.6

0.4 0.4
CB54 CB50
FLUENT
0.3 0.3 FLUENT
Measured
Measured

Fs (g/m/s)
Fs (g/m/s)

0.2 0.2

0.1 0.1

x (m) x (m)
0 0
0 0.2 0.4 0.6 0 0.2 0.4 0.6

(a) Coarse Sand


0.4 0.4
FB66 FB58

FLUENT FLUENT
0.3 0.3
Measured Measured
Fs (g/m/s)

Fs (g/m/s)

0.2 0.2

0.1 0.1

x (m) x (m)
0 0
0 0.2 0.4 0.6 0 0.2 0.4 0.6

0.4
FB46

0.3 FLUENT
Measured
Fs (g/m/s)

0.2

0.1

x (m)
0
0 0.2 0.4 0.6

(b) Fine Sand

Figure 6.19: Comparison of CFD predicted and measured longitudinal deposition


profile (g/m/s) of Li (2006)s horizontal sediment-laden buoyant jet experiments.

173
Jet centerline sediment concentration Jet centerline tracer concentration
Cs/Cs0 Cc/C0
0.35 1 0.35 1
0.9 0.9
0.3 0.3
0.8 0.8
0.25 0.7 0.25 0.7
0.6 0.6
0.2 0.5 0.2 0.5
0.4 0.4
0.15 0.3 0.15 0.3
0.1 0.2 0.1 0.2
0.1 0.1
0.05 0.05 0.05 0.05
0.01 0.01
0 0

-0.05 -0.05
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6
(a) CB66: u0 = 0.65m/s, Fr = 19.5, deq = 166m, ws = 1.95cm/s
Jet centerline sediment concentration Jet centerline tracer concentration
Cs/Cs0 Cc/C0
0.35 1 0.35 1
0.9 0.9
0.3 0.3
0.8 0.8
0.25 0.7 0.25 0.7
0.6 0.6
0.2 0.5 0.2 0.5
0.4 0.4
0.15 0.3 0.15 0.3
0.1 0.2 0.1 0.2
0.1 0.1
0.05 0.05 0.05 0.05
0.01 0.01
0 0

-0.05 -0.05
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6
(b) CB62: u0 = 0.61m/s, Fr = 18.8, deq = 166m, ws = 1.95cm/s
Jet centerline sediment concentration Jet centerline tracer concentration
Cs/Cs0 Cc/C0
0.35 1 0.35 1
0.9 0.9
0.3 0.3
0.8 0.8
0.25 0.7 0.25 0.7
0.6 0.6
0.2 0.5 0.2 0.5
0.4 0.4
0.15 0.3 0.15 0.3
0.1 0.2 0.1 0.2
0.1 0.1
0.05 0.05 0.05 0.05
0.01 0.01
0 0

-0.05 -0.05
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6
(c) CB54: u0 = 0.53m/s, Fr = 16.4, deq = 166m, ws = 1.95cm/s
Jet centerline sediment concentration Jet centerline tracer concentration
Cs/Cs0 Cc/C0
0.35 1 0.35 1
0.9 0.9
0.3 0.3
0.8 0.8
0.25 0.7 0.25 0.7
0.6 0.6
0.2 0.5 0.2 0.5
0.4 0.4
0.15 0.3 0.15 0.3
0.1 0.2 0.1 0.2
0.1 0.1
0.05 0.05 0.05 0.05
0.01 0.01
0 0

-0.05 -0.05
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6
(d) CB50: u0 = 0.49m/s, Fr = 14.9, deq = 166m, ws = 1.95cm/s

Figure 6.20: CFD predicted sediment concentration and conservative tracer con-
centration for buoyant jets, coarse sand experiments. The dashed line is the jet
trajectory and top-hat boundary.

174
Jet centerline sediment concentration Jet centerline tracer concentration
Cs/Cs0 Cc/C0
0.35 1 0.35 1
0.9 0.9
0.3 0.3
0.8 0.8
0.25 0.7 0.25 0.7
0.6 0.6
0.2 0.5 0.2 0.5
0.4 0.4
0.15 0.3 0.15 0.3
0.1 0.2 0.1 0.2
0.1 0.1
0.05 0.05 0.05 0.05
0.01 0.01
0 0

-0.05 -0.05
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6
(a) FB66: u0 = 0.65m/s, Fr = 20.8, deq = 133m, ws = 1.36cm/s
Jet centerline sediment concentration Jet centerline tracer concentration
Cs/Cs0 Cc/C0
0.35 1 0.35 1
0.9 0.9
0.3 0.3
0.8 0.8
0.25 0.7 0.25 0.7
0.6 0.6
0.2 0.5 0.2 0.5
0.4 0.4
0.15 0.3 0.15 0.3
0.1 0.2 0.1 0.2
0.1 0.1
0.05 0.05 0.05 0.05
0.01 0.01
0 0

-0.05 -0.05
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6
(b) FB58: u0 = 0.61m/s, Fr = 14.5, deq = 133m, ws = 1.36cm/s
Jet centerline sediment concentration Jet centerline tracer concentration
Cs/Cs0 Cc/C0
0.35 1 0.35 1
0.9 0.9
0.3 0.3
0.8 0.8
0.25 0.7 0.25 0.7
0.6 0.6
0.2 0.5 0.2 0.5
0.4 0.4
0.15 0.3 0.15 0.3
0.1 0.2 0.1 0.2
0.1 0.1
0.05 0.05 0.05 0.05
0.01 0.01
0 0

-0.05 -0.05
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6
(c) FB46: u0 = 0.53m/s, Fr = 14.5, deq = 133m, ws = 1.36cm/s

Figure 6.21: CFD predicted sediment concentration and conservative tracer con-
centration for buoyant jets, fine sand experiments. The dashed line is the jet
trajectory and top-hat boundary.

175
6.5.3 Particle model prediction
In this section the particle tracking model prediction is discussed in detail with
comparison with data. The simplified model is used for all 31 cases with sand
and glass particles (Tables 6.4-6.6). The Full Model is used for the 4 cases with
plastic particles (Tables 6.7 and 6.8). For each case, the flow field is firstly
predicted using the methods described in Section 6.2. Np = 10000 particles are
used for each simulation. The particles are initialized at the end of the zone of
flow establishment (length deduced before the simulation of JETLAG). For Li
(2006) and Lee (2010)s experiments using ethanol-water mixture as source fluid,
the local fluid density and viscosity is estimated based on the relations shown in
Appendix C to predict the particle velocity locally. The longitudinal deposition
profile is determined by summing all particle masses at the y-direction under
intervals of x = 0.04m.

Experiments of Li (2006)

The seven experiments of Li (2006) are carried out with jet initial densimetric
Froude number F r > 14. For the four experiments using coarse sand particles
(Fig. 6.22), the lm s (0.17-0.26m) are all smaller than 3ls (0.25-0.33m), indicating
the bottom deposition pattern is dominated by falling out from the jet region.
The deposition from spreading current is of minor importance. Similar to a
horizontal momentum jet, the higher the jet velocity, the further the location of
the maximum deposition rate from the jet nozzle and the smaller the maximum
deposition rate. The model predictions compare well with experimental data.
For the three experiments using fine sands (Fig. 6.23), their lm s (0.26-0.49m)
are all larger than 3ls (0.25-0.35m), indicated that a substantial portion of sed-
iment is transported to the surface spreading current by the curving upward
jet flow. A double peak deposition structure is formed. The numerical particle
model well predicts the location of the second peak and compares much better
than the CFD prediction (Fig. 6.19).
Comparing the 2D particle deposition patterns (Fig. 6.24) of the coarse and
fine sand experiments, it can be seen that in both experiment there is a narrow
zone of high particle density closer to the jet nozzle. Its transverse extent spreads
up further downstream. This zone refers to the jet fall out and similar to those
observed in horizontal momentum jets (Chapter 5). The deposition spreads up
dramatically as a circular pattern. This is the deposition from the spreading
layer. For coarse sand experiments, the circular deposition pattern is hardly
visible as only a small portion is brought up to the spreading current. For fine
sand experiment, the circular spread is larger and consists of more particles,
indicates the dominance of the fall out from spreading current.
Fig. 6.25 shows the several trajectories of particles in the fine sediment case
FB66, indicating the complex dynamics of particles in buoyant jets with surface
impingement. Particles can fall out directly from the jet before it bends up
(Trajectory i). They can also rise up to the surface spreading layer, fall out and
move along the jet boundary due to the entrainment induced current, but without

176
being entrained back into the jet (Trajectory ii). As the most complicated,
particles can move within the impingement zone and re-entrained several times
before they deposit in the bottom (Trajectory iii).

Experiments of Lee (2010)

Six experiments were carried out using glass particles S199 and S153, with rela-
tively low jet densimetric Froude number (F r=10-14, Table 6.5). The buoyant
jets bent up quickly after a short distance of horizontal travel. For S199 ex-
periments (Fig. 6.26), both measured and predicted deposition pattern shows
a significant one-peak structure due to fall out from the jet, as lm < 3ls for
all cases (lm = 0.11 0.14m, 3ls = 0.17 0.25m). For the S153 experiments
(Fig. 6.27), their lm and 3ls are quite similar, both 0.16-0.20m. A significant
portion of particles is brought to the surface current before they settle out from
the jet region. The close lm and 3ls (Table 6.5) indicates that the first and second
peak merges together to form a flat plateau. For both cases, the depositions are
well predicted.
Another 18 experiments were carried out using the G215, G180 and G115
spherical glass particles which are the same as those in present study. Six ex-
periments were carried out for each type of particles (Table 6.6). Two sets of
flow rate - 50 L/s and 90L/s are tested for each type of particles. In each flow
rate, the densimetric Froude numbers are controlled by varying the density of
the source fluid so that different trajectories are obtained.
For G215 experiments, the particles have a high settling velocity of about
2.65cm/s. The lm (0.10m, 0.18m) are all much smaller than 3ls (0.17-0.47m).
Most sediment is settled from the jet region, which insignificant portion brought
up to the spreading current. The bottom deposition is well predicted for these
momentum jet-like cases.
For G180 experiments, the deposition profiles are also jet-fall-out dominated
as lm (0.13m, 0.23m) is in general smaller than ls (0.17-0.47m). For some cases
with low F r, the second peak is more significant, like G180J90Fr17, G180B50Fr12
and G180B50Fr10. The predictions compares very well with the measurement.
The G115 experiments are the most challenging for prediction, as the most
particles are transported to the spreading current and their dynamics are influ-
enced by the external current. For these experiments, lm (0.26m and 0.47m) is
all close to, or greater than 3ls . This results in a significant portion of sediment
brought to the surface current and the second peak. From the measurement re-
sults it is observed that the deposition of G115B50Fr10 and G115B50Fr12 cases
of Q = 50L/h are completely due to the surface spreading current fall out. The
particle model predicts the observed single peak very well, despite the predicted
profile has a large spread and smaller peak.
Significant discrepancies are also seen in the high flow cases of G115B50Fr22
and G115B50Fr28. The predicted second peak is much higher and closer to the
jet nozzle. This is attributed to two reasons: 1) the spreading current prediction
ignored the horizontal momentum, which is is highest for the high flow rate of

177
0.5 0.5
Fs (g/m/s) CB66 Fs (g/m/s) CB62
0.4 0.4

0.3 Model Pred. 0.3


Meas.
0.2 0.2

0.1 0.1
x (m) x (m)
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

(a) CB66: u0 = 0.65m/s, Fr = 19.5 (b) CB62: u0 = 0.61m/s, Fr = 18.8


deq = 166m, ws = 1.95cm/s deq = 166m, ws = 1.95cm/s

0.5 0.5
Fs (g/m/s) CB54 F (g/m/s) CB50
0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1
x (m) x (m)
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

(c) CB54: u0 = 0.53m/s, Fr = 16.4 (d) CB50: u0 = 0.49m/s, Fr = 14.9


deq = 166m, ws = 1.95cm/s deq = 166m, ws = 1.95cm/s

Figure 6.22: Comparison of particle tracking model predicted and measured


longitudinal deposition profile (g/m/s) of Li (2006)s horizontal sediment-laden
buoyant jet experiments, Coarse Sand.

0.5 0.5
Fs (g/m/s) FB66 Fs (g/m/s) FB58
Predicted
0.4 0.4
Measured
0.3 0.3

0.2 0.2

0.1 0.1

x (m) x (m)
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

(a) FB66: u0 = 0.65m/s, Fr = 20.8 (b) FB58: u0 = 0.61m/s, Fr = 14.5


deq = 133m, ws = 1.36cm/s deq = 133m, ws = 1.36cm/s

0.5
FB46
Fs (g/m/s)
0.4

0.3

0.2

0.1
x (m)
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

(c) FB46: u0 = 0.53m/s, Fr = 14.5


deq = 133m, ws = 1.36cm/s

Figure 6.23: Comparison of particle tracking model predicted and measured


longitudinal deposition profile (g/m/s) of Li (2006)s horizontal sediment-laden
buoyant jet experiments, Fine Sand.

178
0.4
y (m) CB66

0.3

0.2

0.1

-0.1

-0.2

-0.3

x (m)
-0.4
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
(a) CB66: u0 = 0.65m/s, F r = 19.5, deq = 166m , ws =
1.95cm/s

0.4
y (m) FB58

0.3

0.2

0.1

-0.1

-0.2

-0.3

x (m)
-0.4
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
(b) FB58: u0 = 0.61m/s, F r = 14.5, deq = 133m , ws =
1.36cm/s

Figure 6.24: Predicted 2D deposition profiles for (a) CB66 and (b) FB58.

179
0.38
z (m)
0.33
0.28 (iii)
0.23
0.18
0.13
(ii)
0.08
0.03
-0.02 (i) x (m)
-0.07
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

Figure 6.25: Examples of particle trajectory in case FB66 with fine sediment. The
dash-dotted line is the jet centerline; dashed lines are the jet top-hat boundaries

0.7 0.7
Fs (g/m/s) S199B46Fr9 Fs (g/m/s) S199B58Fr12
0.6 0.6
0.5 0.5
0.4 Predicted 0.4
Measured
0.3 0.3
0.2 0.2
0.1 0.1
x (m) x (m)
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

(a) S199B46Fr9: u0 = 0.45m/s, Fr = 9.9 (b) S199B58Fr12: u0 = 0.56m/s, Fr = 12.2


d50 = 199m, ws = 2.20cm/s d50 = 199m, ws = 2.24cm/s

0.7
Fs (g/m/s) S199B50Fr14
0.6
0.5
0.4
0.3
0.2
0.1
x (m)
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

(c) S199B50Fr14: u0 = 0.49m/s, Fr = 14.6


d50 = 199m, ws = 2.21cm/s

Figure 6.26: Comparison of particle tracking model predicted and measured


longitudinal deposition profile (g/m/s) of Lee (2010)s horizontal sediment-laden
buoyant jet experiments, S199 particles.

180
0.5 0.4
Fs (g/m/s) S153B46Fr9 Fs (g/m/s) S153B52Fr11
Predicted
0.4
Measured 0.3
0.3
0.2
0.2
0.1
0.1

x (m) x (m)
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

(a) S153B46Fr9: u0 = 0.45m/s, Fr = 9.8 (b) S153B52Fr11: u0 = 0.51m/s, Fr = 11.2


d50 = 153m, ws = 1.50cm/s d50 = 153m, ws = 1.50cm/s

0.4
Fs (g/m/s) S153B58Fr12

0.3

0.2

0.1

x (m)
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

(c) S153B58Fr12: u0 = 0.56m/s, Fr = 12.2


d50 = 153m, ws = 1.50cm/s

Figure 6.27: Comparison of particle tracking model predicted and measured


longitudinal deposition profile (g/m/s) of Lee (2010)s horizontal sediment-laden
buoyant jet experiments, S153 particles.

Q = 90L/s (u0 = 0.88m/s); 2) with such high initial velocity, the jet trajectory
covers almost half of the tank (Fig. 6.31a), compared with a low flow buoyant jet
(e.g. u0 = 0.49m/s, Fig. 6.31b). The entrainment induced flow and the spreading
current in this limited-sized tank may be more complicated as the influence of
boundary is significant. Except for these two cases, the particle model predicts
the deposition for horizontal buoyant jets well.

181
1 1
F (g/m/s) G215B90Fr27 Fs (g/m/s) G215B90Fr22
0.8 0.8

0.6 Predicted 0.6


Measured
0.4 0.4

0.2 0.2
x (m) x (m)
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

(a) G215B90Fr27: u0 = 0.88m/s, Fr = 27.9 (b) G215B90Fr22: u0 = 0.88m/s, Fr = 22.5


d50 = 215m, ws = 2.70cm/s d50 = 215m, ws = 2.70cm/s

1 1
F (g/m/s) G215B90Fr17 Fs (g/m/s) G215B50Fr15
0.8 0.8
Predicted
0.6 0.6 Measured

0.4 0.4

0.2 0.2
x (m) x (m)
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

(c) G215B90Fr17: u0 = 0.88m/s, Fr = 18.1 (d) G215B50Fr15: u0 = 0.49m/s, Fr = 15.5


d50 = 215m, ws = 2.70cm/s d50 = 215m, ws = 2.70cm/s

1 1
Fs (g/m/s) G215B50Fr12 Fs (g/m/s) G215B50Fr10
0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2
x (m) x (m)
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

(e) G215B50Fr12: u0 = 0.49m/s, Fr = 12.4 (f) G215B50Fr10: u0 = 0.49m/s, Fr = 10.1


d50 = 215m, ws = 2.70cm/s d50 = 215m, ws = 2.70cm/s

Figure 6.28: Comparison of particle tracking model predicted and measured


longitudinal deposition profile (g/m/s) of Lee (2010)s horizontal sediment-laden
buoyant jet experiments, G215 particles.

182
0.5 0.5
Fs (g/m/s) G180B90Fr27 Fs (g/m/s) G180B90Fr22
0.4 0.4
Predicted
0.3 Measured 0.3

0.2 0.2

0.1 0.1
x (m) x (m)
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

(a) G180B90Fr27: u0 = 0.88m/s, Fr = 28.3 (b) G180B90Fr22: u0 = 0.88m/s, Fr = 22.7


d50 = 180m, ws = 2.06cm/s d50 = 180m, ws = 2.06cm/s

0.5 0.5
Fs (g/m/s) G180B90Fr17 Fs (g/m/s) G180B50Fr15
0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1
x (m) x (m)
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

(c) G180B90Fr17: u0 = 0.88m/s, Fr = 18.3 (d) G180B50Fr15: u0 = 0.49m/s, Fr = 15.5


d50 = 180m, ws = 2.06cm/s d50 = 180m, ws = 2.06cm/s

0.5 0.5
Fs (g/m/s) G180B50Fr12 Fs (g/m/s) G180B50Fr10
0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1
x (m) x (m)
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

(e) G180B50Fr12: u0 = 0.49m/s, Fr = 12.4 (f) G180B50Fr10: u0 = 0.49m/s, Fr = 10.0


d50 = 180m, ws = 2.06cm/s d50 = 180m, ws = 2.06cm/s

Figure 6.29: Comparison of particle tracking model predicted and measured


longitudinal deposition profile (g/m/s) of Lee (2010)s horizontal sediment-laden
buoyant jet experiments, G180 particles.

183
0.2 0.2
Fs (g/m/s) G115B90Fr28 Fs (g/m/s) G115B90Fr22

Predicted
Measured
0.1 0.1

x (m)
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 x (m)0.8 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

(a) G115B90Fr27: u0 = 0.88m/s, Fr = 28.3 (b) G115B90Fr22: u0 = 0.88m/s, Fr = 22.7


d50 = 115m, ws = 1.0cm/s d50 = 115m, ws = 1.0cm/s

0.2 0.2
Fs (g/m/s) G115B90Fr17 Fs (g/m/s) G115B50Fr15

0.1 0.1

x (m) x (m)
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

(c) G115B90Fr17: u0 = 0.88m/s, Fr = 18.3 (d) G115B50Fr15: u0 = 0.49m/s, Fr = 15.8


d50 = 115m, ws = 1.0cm/s d50 = 115m, ws = 1.0cm/s

0.2 0.2
Fs (g/m/s) G115B50Fr12 Fs (g/m/s) G115B50Fr10

0.1 0.1

x (m) x (m)
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

(e) G115B50Fr12: u0 = 0.49m/s, Fr = 12.5 (f) G115B50Fr10: u0 = 0.49m/s, Fr = 10.2


d50 = 115m, ws = 1.0cm/s d50 = 115m, ws = 1.0cm/s

Figure 6.30: Comparison of particle tracking model predicted and measured


longitudinal deposition profile (g/m/s) of Lee (2010)s horizontal sediment-laden
buoyant jet experiments, G115 particles.

0.4 0.4
z (m) z (m)
0.3 0.3

0.2 0.2

0.1 0.1

0 0
x (m) x (m)
-0.1 -0.1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

(a) Qj = 90L/h, u0 = 0.88m/s, F r = 28.3 (b) Qj = 50L/h, u0 = 0.49m/s, F r = 10.2

Figure 6.31: JETLAG predicted jet trajectories and boundary profiles.

184
6.5.4 Plastic particle experiments: Lee (2010), Cuthbert-
son and Davies (2008)
Horizontal sediment-laden buoyant jet experiments using plastic particles are
used to validate the full model on bottom deposition prediction. Lee (2010)
reported two buoyant jet experiment using granular plastic particles of density
p = 1.16 g/cm3 (Table 6.7) carried out using similar experimental apparatus as
in this study.
Cuthbertson and Davies (2008) carried out plastic particle-laden buoyant
jet experiments in a water channel of length 10m, width 1m and depth 0.75m.
The jet nozzle of 11.5mm internal diameter is located at the mid-way of the
channel and 15cm above the channel bottom, discharging freshwater to saline
ambient. 43 experiments in stagnant ambient and 32 experiments in coflowing
ambient are reported. Their experiments covers a wide range of jet discharge
(Q0 = 3 10L/min) and buoyancy conditions (F r = 7.2 35), with two types
of plastic particles used. Photographic technique is used to measure bottom
deposition. In this study, two of the experiments using POLY1 particles (p = 1.5
g/cm3 , d = 550m ) are used for model validation (Table 6.8).
The comparison of model prediction and measurement of Lees experiments
is shown in Fig. 6.32. The comparison is very good for the Full Model predic-
tion. Consistent with the previous conclusion, the Simplified Model prediction
is less satisfactory for these lighter plastic particles. Fig. 6.33 shows the model
prediction with measured 1D and 2D deposition data by Cuthbertson and Davies
(2008). The difference in the Full and Simplified Models can be observed clearly
in these cases. Due to the particle inertia effect, generalization of the previous
experimental results obtained by dimension analysis using particles of different
properties has to be viewed with caution. The predicted 2D deposition profiles
compare qualitatively well with the photographic measurement.

Table 6.7: Horizontal buoyant jet experiments of Lee (2010), plastic particles
(d = 621m , p = 1.16 g/cm3 ).
Case u0 C0 ws0 wsa a 0 Fr lm 3ls
(m/s) (g/L) (cm/s) (cm/s) (kg/m3 ) (kg/m3 ) (m) (m)
PB50Fr12 0.4912 1.57 1.87 2.09 996.0 964.7 10.9 0.1397 0.1935
PB66Fr15 0.6484 1.36 1.83 2.09 996.8 966.4 14.6 0.1884 0.2593

Jet diameter = 6mm


1/2
lm = M0 /ws = (/4)1/2 u0 D/ws
3/4 1/2
ls = M0 /B0 = (/4)1/4 F rD

185
Table 6.8: Horizontal buoyant jet experiments of Cuthbertson and Davies (2008),
plastic particles (d = 550m , p = 1.5 g/cm3 ).
Case u0 ws a 0 Fr lm 3ls
(m/s) (cm/s) (kg/m3 ) (kg/m3 ) (m) (m)
341 1.041 2.97 1020.0 1000.0 22.1 0.357 0.718
365 1.295 2.97 1040.0 1000.0 19.7 0.444 0.640

Jet diameter = 11.5mm


1/2
lm = M0 /ws = (/4)1/2 u0 D/ws
3/4 1/2
ls = M0 /B0 = (/4)1/4 F rD

0.3 0.3
F (g/m/s) PB50Fr12 F (g/m/s) PB66Fr15
Full Model Full Model
0.2 Simp. Model 0.2 Simp. Model
Meas. Meas.

0.1 0.1

x (m) x (m)
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

(a) PB50Fr12: u0 = 0.49m/s, Fr = 10.9 (b) PB66Fr15: u0 = 0.65m/s, Fr = 14.6


d50 = 621m, ws = 2.09cm/s d50 = 621m, ws = 2.09cm/s

Figure 6.32: Comparison of particle tracking model predicted and measured


longitudinal deposition profile (g/m/s) of Lee (2010)s horizontal sediment-laden
buoyant jet experiments, plastic particles (p = 1.16 g/cm3 ).

186
0.4 0.4
y (m) 341

0.2 0.2

0 0

-0.2 -0.2

x (m)
-0.4 -0.4
0 0.4 0.8 1.2 0 0.4 0.8 1.2

(i) Observed (i) Observed


0.4 0.4
y (m) 341 y (m) 365

0.2 0.2

0 0

-0.2 -0.2

x (m) x (m)
-0.4 -0.4
0 0.4 0.8 1.2 0 0.4 0.8 1.2

(ii) Predicted (ii) Predicted


4 4
-1 341 -1
3.5 F/Q0 C0 (m ) 3.5 F/Q0 C0 (m ) Full Model 365
3 Full Model 3 Simp. Model
2.5 Simp. Model 2.5 Meas.
2 Measured 2
1.5 1.5
1 1
0.5 0.5
x (m) x (m)
0 0
0 0.2 0.4 0.6 0.8 1 1.2 0 0.2 0.4 0.6 0.8 1 1.2

(iii) 1D deposition profile (iii) 1D deposition profile

(a) Case 341: u0 = 1.04m/s, Fr = 22.1, ws = (b) Case 365: u0 = 1.30m/s, Fr = 19.7, ws =
2.97cm/s 2.97cm/s

Figure 6.33: Comparison of model predicted and measured 1D and 2D deposition


profiles Cuthbertson and Davies (2008)s horizontal sediment-laden buoyant jet
experiments, plastic particles (p = 1.5 g/cm3 ). The 1D deposition profiles are
normalized by the sediment mass flux at the jet nozzle, in the unit of m1 .

187
6.6 Summary
In this chapter, a general model for the prediction of sediment deposition from
buoyant jets in stagnant water is developed. The three flow regimes of a buoyant
jet impinging a free surface, namely, the mean jet velocity, external entrainment
current and surface spreading current, can be predicted using semi-analytical
models:

1. The integral Lagrangian model JETLAG is used to predict the trajectory


and mean flow velocities of buoyant jet. Turbulent fluctuations are pre-
dicted using the CFD solutions of turbulent intensity and dissipation rate
with a two-equation turbulence closure model for pure jets and plumes.

2. The external flow is predicted using the point sink approach.

3. The surface spreading current is predicted using a semi-analytical integral


model accounting for interfacial shear and validated with CFD prediction.

The model is successfully applied to predict the deposition from vertical buoy-
ant jets with two distinct regimes of particle fall out: (1) fall out from jet margin,
and (2) fall out from spreading current. Model prediction is well compared to
experimental data in past studies and supports the observation that sediment
deposits the closest to the jet nozzle in the jet margin fall out from plumes, as
a plume have the highest entrainment velocity. On the other hand, sediment
fall out from surface current has the greatest spread. Sensitivity analysis shows
that both Full Model and Simplified Model works well for these independent
experiments using polystyrene and silicon carbide particles, and the Basset force
is negligible.
The model is further validated with extensive data of horizontal sediment-
laden buoyant jets, including sand, glass and plastic particles (specific gravity
s = 1.16 2.65). The two modes of deposition: fall out from jet lower boundary
and fall out from spreading current, are very well predicted.

188
Chapter 7

Conclusion

7.1 Summary of this study

Sediment or particle-laden buoyant jets are common in natural environment and


engineering applications. The objective of the present study is to develop a gen-
eral model to predict the sediment mixing and deposition of dilute sediment-laden
buoyant jets in arbitrary inclination in stagnant ambient. This work reports a
theoretical, numerical and experimental investigation on the mixing and sedi-
mentation of particulate matters from sediment-laden buoyant jets.

(i) Particle tracking model and autocorrelation function


A particle tracking model is developed for the motion of sediment in turbu-
lence, using a velocity autocorrelation function that describes the trapping and
loitering effect of particles in turbulence. Two particle tracking models are devel-
oped to predict the movement of sediment particles in turbulent flows. The Full
Model solves the governing equation of particle motion numerically. The Sim-
plified Model assumes the particle velocity is the sum of the fluid velocity and
the stillwater settling velocity. Measurement of single phase jet velocity is carried
out for the experimental support of the autocorrelation function. The positive
correlation between the radial (vertical) velocity magnitude and the absolute
change in velocity supports the assumption in the autocorrelation function.
The particle tracking model is validated with results in independent stud-
ies, including (1) analytical solution and experimental measurement of parti-
cle motion in stagnant fluid; (2) experimental measurement on particle settling
in homogeneous vertically oscillating fluid field; (3) experimental measurement
on particle settling in oscillating grid turbulence and (4) vertically downward
sediment-laden jets.

(ii) Horizontal sediment-laden momentum jets


The mixing and deposition of sediment-laden horizontal momentum jets are
studied using laboratory experiments and computational fluid dynamics mod-
elling (CFD). Laboratory experiments and CFD modelling shows that there is

189
a significant settling velocity reduction up to about 25-35% under the influence
of jet turbulence, which is dependent on the RMS turbulent velocity to stillwa-
ter settling velocity ratio /ws and the particle properties. Using a relationship
for estimating the reduction in settling velocity locally based on the /ws ra-
tio, the Eulerian CFD model predicted bottom deposition of sand and glass
particle-laden jets are significantly improved as compared to using directly the
stillwater settling velocity. However, this CFD approach necessitates an ad hoc
adjustment/reduction of particle settling velocity and lacks generality.
A stochastic particle tracking model is developed to predict the bottom depo-
sition and sediment concentration in horizontal sediment-laden momentum jets.
Analytical solutions for a pure jet is used. Turbulent velocity fluctuations and
turbulent time scales are modelled with best-fitted self-similar RMS turbulent
velocity and dissipation rate profiles derived from CFD solution of a pure jet.
The model has been fully validated on sediment-laden jet experimental data,
with initial jet velocity u0 = 0.3 0.9m/s and covering a wide range of par-
ticle properties, including sand, glass and plastic particle (p =1.1-2.6 g/cm3 ,
d = 115 716m , ws = 1.0 2.7cm/s)
Model predictions are in excellent agreement with experimental data. The
particle tracking model is superior to a 3D Eulerian CFD model as it does not
require any a priori empirical adjustment to the stillwater settling velocity. Sup-
ported by experimental data, a sensitivity study shows that the computationally
demanding Basset history force is negligible for all kind of particles used in
present sediment-laden jet experiments.

(iii) A general model for arbitrarily inclined sediment-laden buoyant


jets
A general model for the prediction of sediment deposition from buoyant jets
in stagnant water is developed. The three flow regimes of a buoyant jet impinging
a free surface, namely, turbulent jet flow, external entrainment flow and surface
spreading current, can be predicted using semi-analytical models. The integral
Lagrangian model JETLAG is used to predict the trajectory and mean flow
velocities of the buoyant jet. Turbulent fluctuations are predicted using the
CFD derived profiles of turbulent intensity and dissipation rate with a two-
equation turbulence closure model. The external flow is predicted using a recently
developed point sink approach. The surface spreading current is predicted using
a semi-analytical integral approach and validated with CFD prediction.
The model is successfully applied to predict the deposition from vertical buoy-
ant jets with two distinct regimes of particle fall out: (1) fall out from jet margin,
and (2) fall out from spreading current. Model prediction is in good agreement
with past experimental data. Sensitivity analysis shows that both Full Model and
Simplified Model works well for these independent experiments using polystyrene
(specific gravity s = 1.04) and silicon carbide particles (s = 3.2), and the Basset
force is negligible.
The model is further validated with horizontal sediment-laden buoyant jets
with extensive experimental data, including sand, glass and plastic particles (s =

190
1.16 2.65). The two modes of deposition: fall out from jet lower boundary and
fall out from spreading current, are very well predicted.

(iv) Overall contribution


This is the first time that a general model for predicting the mixing and
deposition of sediment-laden buoyant jets in arbitrary inclination is successfully
developed and extensively validated with experimental data over a wide range
of jet conditions and particle properties. The turbulence-particle interaction -
trapping and loitering of particles in turbulent eddies in a sediment-laden jet
are supported by experimental observation and numerical modelling for the first
time. This modelling framework is beneficial to many types of geophysical and
engineering problems, including volcanic eruption, hydrothermal vents, sediment
disposal, sewage treatment and discharge and many other industrial applications.

7.2 Recommendations for future work


This work has made significant progress to the modelling and prediction of
sediment-laden jets, yet the topic still has many aspects to be explored. The
following future works are recommended:

1. The present model can be extended to sediment-laden jets in a wide range of


ambient conditions, including ambient current and stratification. Further
laboratory and numerical modelling studies are required to understand the
physics of sediment-laden jets in a cross-flowing and/or stratified ambient
and to validate the model.

2. A dynamic coupling of the near-field JETLAG model with a 3D shallow


water circulation has been developed by Choi and Lee (2007) and validated
with experimental and field data. The present particle model can be in-
corporated in this coupled near-far field system to study the transport of
sediment from sewage outfalls to sensitive receivers seamlessly.

3. Further understanding on the turbulent structure of other turbulent flows


(open channel flow, atmospheric turbulent flow, stratified flow) is required
to support the applicability of the autocorrelation function and understand
the physics of loitering and trapping of particles in turbulent eddies.

4. The effect of sedimentation from outfall plumes on the marine environment


has to be studied and quantified with laboratory experiment and field mea-
surement. The field application of the present model for the design and
assessment of submarine outfalls has to be explored.

191
7.3 Application examples
Example 1 - Vertical sediment-laden jet
In many small communities or developing countries, the lack of capital allows
only for preliminary sewage treatment. This example shows the impact assess-
ment of sewage sludge particles to the recreational and fish cultural activities at
the water surface of an outfall discharging vertically. The outfall is situated at
a depth of 20m. Screened effluent is discharged from the 0.1m diameter outfall
in the velocity of 0.5 m/s (Q0 = 340m3 /d) with density of freshwater of 1000
kg/m3 . The ambient sea water density is 1020 kg/m3 . The suspended solid
concentration is 0.2 g/L, consists of sand particles of median diameter 200 m .
q
The jet densimetric Froude number F r = u0 / gD(a 0 )/a can be deter-
mined as 3.6. The particle settling velocity is determined using Soulsby (1997)s
equation as ws = 2.5 cm/s. The model simulated the situation for 1 hour. Fig. 7.1
shows the sequential 3D depiction of the particle laden jet. It can be observed
that the particles reaches the surface in about 3 min, and transported later-
ally into the spreading current. Particles gradually fall out from the spreading
current, forming a particle cloud moving down the water depth. The falling sed-
iment from the spreading current covers a much larger radial extent compared
with the original buoyant jet. The sediment veil gradually converges towards
the bottom, showing the effect of re-entrainment. Fig. 7.2 shows the predicted
bottom deposition profile. Because of the external entrainment flow, the radial
extent of deposition is about 6m, while particle in the spreading current reaches
a radius of 8m.

192
(a) 3 min (b) 6 min

(c) 12 min (d) 30 min

Figure 7.1: Visualization of a vertical sediment-laden jet discharging upwards -


Example 1. The box is for visualization only. Free lateral boundary is used in
the model.

193
2
Fs (tonnes/m /yr) z (m)
1 20
Spreading current 18
0.8 16
Jet boundary
14
0.6 12
10
0.4 8
Sediment 6
deposition
0.2 4
2
0 0
0 2 4 6 8 r (m) 10

Figure 7.2: Predicted deposition profile of vertical sediment-laden jet discharging


upwards - Example 1. The jet and spreading current profiles are also indicated.

Example 2 - Sewage discharge in form of horizontal sediment buoyant


jet
This example illustrates the assessment of sediment deposition from a single-
jet outfall discharging preliminarily treated domestic sewage in southern Hong
Kong Island - the Wah Fu outfall. Stagnant and unstratified ambient condition
is assumed. The outfall has 14 identical jets spaced large enough to avoid jet
interaction. The outfall discharges a daily averaged flow of 0.088 m3 /s. Ambient
sea water density a is assumed to be 1018 kg/m3 and sewage density 0 is
assumed as 1004 kg/m3 due to the use of seawater for toilet flushing in Hong
Kong. For each jet, the following configuration is assumed.

Jet discharge Q0 = 0.00626 m3 /s

Jet diameter D = 0.1 m

Jet velocity u0 = 0.797 m/s

Jet height from the sea bottom = 2 m

Depth of jet = 7 m

Vertical angle = 20

The initial sediment concentration is 200 mg/L with median diameter of 300
m composed of mainly natural sand of density p = 2650 kg/m3 . About 560
tonnes of sewage sludges per year are discharged to the marine environment
through the outfall. Here the deposition profile of one jet is predicted.
q
The jet densimetric Froude number F r = u0 / gD(a 0 )/a can be deter-
mined as 6.86. The momentum-buoyancy length scale ls = (/4)1/4 F rD = 0.65
m. The particle settling velocity is determined using Soulsby (1997)s equation
as ws = 4.27 cm/s. The momentum-settling length scale lm = (/4)1/2 u0 D/ws =

194
1.65 m. The predicted jet trajectory is shown in Fig. 7.3. Since lm 3ls for this
case, both fall out from jet margin and fall out from the spreading current are
contributing similar portions to the deposition profile. Using the particle model,
the bottom deposition profile can be obtained (Fig. 7.4). The deposition profile
consists of a narrow concentrated zone corresponding to fall out from the jet
margin, and a spread-out circular region corresponding to fall out from surface
current. This gives a general idea on how the sediment forms a sludge bank
close to the outfall and its potential impact to the marine water quality. In re-
ality, the ambient current would also affect the sediment transport and erosion.
Particle size distribution and flocculation are also complicating factors.
7
(m)
6

-1
(m)
-2
0 1 2 3 4 5 6 7

(a) Predicted jet trajectory (b) Observed jet trajectory

Figure 7.3: (a) The predicted jet trajectory of Wah Fu Outfall - prototype. (b)
Experimental image of a 1:11 scale model (Lee and Chu, 2003)

195
Fs (tonnes/m/yr)
30

25

20

15

10

5
lm = 1.65m
0
0 1 2 3 4 5 x (m) 6
(a) Predicted 1D deposition profile (transversely summed)

4
tonnes/m2/yr

3 100
50
20
2 10
5
2
1 1
y (m)

-1

-2

-3

-4
0 2 4 6 8
x (m)
(b) Predicted 2D deposition profile

Figure 7.4: Predicted deposition profiles of horizontal sediment-laden buoyant


jet - Wah Fu Outfall: (a) 1D (in tonnes/m/yr); (b) 2D (in tonnes/m2 /yr)

196
Example 3 - Sediment-laden horizontal buoyant jet in a weak coflowing
current
The present method can be readily extended to predict the sediment deposi-
tion from a horizontal buoyant jet in a weak coflowing current. In a weak current,
the coflow jet is slightly advected but the turbulent motion is similar to that in
stagnant condition. Shear entrainment dominates the increase in jet volumetric
flux. A more general form of the entrainment coefficient (Chapter 6, Eq. 6.5) is
adopted to account for the effect of ambient current (Lee and Chu, 2003).
! 
0.554 sin k 2uk
T = 2 0.057 + (7.1)
Fl2 u + uk

where
u = uk ua cos k (7.2)
is the excess jet velocity and ua is the ambient velocity. The entrainment velocity
ve and local densimetric Froude number Fl is determined based on the excess jet
velocity, i.e.
F u
ve = T u, Fl = r
T,k
a
gbT,k

The stagnant condition equations Eqs. 6.4, 6.5 and 6.6 are recovered when
ua = 0, u = uk . A Gaussian profile is adopted for the streamwise jet excess
velocity. The jet induced irrotational flow field in excess of the ambient velocity
can be predicted using the point sink approach (Lai, 2009; Lai and Lee, 2012a).
The spreading current in excess of the ambient velocity is determined using the
integral model developed in Section 6.2.3. The ambient current is superimposed
on the predicted excess velocity flow field.
Using the Wah Fu outfall as example, the sediment deposition profile is pre-
dicted with an ambient current of ua = 0.03 m/s in the x-direction. A length scale
delineating the asymptotic regions of buoyancy-dominated near field (BDNF)
and buoyancy-dominated far field (BDFF) is defined as (Wood et al. , 1993; Lee
and Chu, 2003)
F0 Q0 g
lb = 3 = 3 0 (7.3)
ua ua
Using the jet parameters in Example 2, lb can be estimated as 31m. Since the
height of rise of the jet is only 7m, the jet is completely within the BDNF region
(z << lb ) in which shear entrainment applies. Vortex entrainment dominates in
the BDFF region (z >> lb ).
The predicted jet trajectory is shown in Fig. 7.5; the jet is advected down-
stream by the ambient current. Predicted sediment deposition profile is dif-
ferent from the stagnant ambient case (Fig. 7.6). The whole profile is shifted
downstream by about 1.5m, which is consistent with the advection of sediment
particles during their settling for a vertical distance of 2m from the jet
  !
z 2m
x = ua = 0.03m/s = 1.4m
ws 0.043m/s

197
Secondly, there is no spreading-out region in the deposition profile, indicating
that sediment particles mainly fall out from the jet margin. Sediment is blown
out of the rising plume by the ambient current, leaving very little sediment
reaching the spreading layer (Fig. 7.5).

7
(m)
6

5 ua = 0.03m/s

-1
(m)
-2
0 2 4 6 8 10

Figure 7.5: Predicted jet trajectory of Wah Fu Outfall in a coflowing ambient


current of 0.03m/s and a visualization of the predicted sediment distribution on
the centerline plane.

198
Fs (tonnes/m/yr)
30

25

ua = 0
20
ua = 0.03m/s

15

10

0
0 2 4 6 8 x (m) 10
(a) Predicted 1D deposition profile (transversely summed)

4
tonnes/m2/yr

3 100
50
20
2 10
5
2
1 1
y (m)

-1

-2

-3

-4
0 2 4 6 8
x (m)
(b) Predicted 2D deposition profile

Figure 7.6: Predicted deposition profiles of horizontal sediment-laden buoyant jet


- Wah Fu Outfall with an ambient current of 0.03m/s: (a) 1D (in tonnes/m/yr);
(b) 2D (in tonnes/m2 /yr)

199
200
Bibliography

[1] Abraham, G. and Eysink, W.D. (1971). Magnitude of interfacial shear in


exchange flow. Journal of Hydraulic Research, 9(2).

[2] Acton, F.S. (1970). Numerical Methods That Work, Harper & Row.

[3] Akar, P.J. and Jirka, G.H. (1994). Buoyant spreading processes in pollu-
tant transport and mixing. Part 1: Lateral spreading with ambient current
advection. Journal of Hydraulic Research, 32(6), 815-831.

[4] Alexander, P. (2004). High order computation of the history term in the
equation of motion for a spherical particle in a fluid. Journal of Scientific
Computing, 21(2), 129-143.

[5] Basset, A.B. (1888). On the Motion of a Sphere in a Viscous Liquid.


Philosophical Transactions of the Royal Society of London A, 179, 43-63

[6] Bleninger, T., Carmer, C. V., Jirka, G., and Neves, M. (2002). Sedimenta-
tion from low concentration particle-laden jets. Proc. 2nd Int. Conf. Marine
Waste Water Discharges, Istanbul, Sept 16-20.

[7] Bombardelli, F.A., Gonzalez, A.E. and Nino, Y.I. (2008). Computation of
the particle Basset force with a fractional-derivative approach., Journal of
Hydraulic Engineering, ASCE, 134(10), 1513-1520.

[8] Brush, Jr.L.M., Ho, H.-W. and Yen, B.-C. (1964). Accelerated motion of
a sphere in a viscous fluid., Journal of Hydraulics Division, Proceedings of
ASCE, 90(HY1), 149-160.

[9] Buhler, J. and Papantoniou, D.A. (2001). On the motion of suspension


thermals and particle swarms. Journal of Hydraulic Research, 39(6), 643-
653.

[10] Bush, J.W.M., Thurber, B.A. and Blanchette, F. (2003). Particle clouds
in homogeneous and stratified environments. Journal of Fluid Mechanics,
489, 29-54.

[11] Cardoso, S. S. S. and Zarrebini, M. (2002). Sedimentation from surface


currents generated by particle-laden jets. Chemical Engineering Science,
57(8), 14251437.

201
[12] Carey, S.N., Sigurdsson, H., and Sparks, R.S.J. (1988). Experimental stud-
ies of particle-laden plumes. Journal of Geophysical Research, 93(B12),
15314-15328.

[13] Cederwall, K. (1968). Hydraulics of marine wastewater disposal. Hy-


draulics Div., Chalmers Institute of Technology, Sewden, Report No.42.

[14] Cheng, N.-S. and Law, A.W.-K. (2001). Measurements of turbulence gener-
ated by oscillating grid. Journal of Hydraulic Engineering, ASCE, 127(3),
201208.

[15] Choi, K.W. and Lee, J.H.W. (2007). Distributed Entrainment Sink Ap-
proach for Modelling Mixing and Transport in the Intermediate Field.
Journal of Hydraulic Engineering, ASCE, 133, 804-815.

[16] Clift, R. and Gauvin, W.H. (1970). The motion of particles in turbulent
gas streams, Proceedings of Chemeca 70, Melbourne and Sydney, 1, 14-28.

[17] Cromey, C.J., Black, K.D., Edwards, A. and Jack, I.A. (1998) Modelling
the deposition and biological effects of organic carbon from marine sewage
discharges. Estuarine, Coastal and Shelf Science, 47, 295-308.

[18] Csanady, G.T. (1963) Turbulent diffusion of heavy particles in the atmo-
sphere. Journal of the Atmospheric Sciences, 20, 201-208.

[19] Csanady, G.T. (1973). Turbulent Diffusion in the Environment. D.Reidel


Publishing Company, Holland

[20] Cuthbertson, A.J.S. and Davies, P.A. (2008) Deposition from Particle-
Laden, Round, Turbulent, Horizontal, Buoyant Jets in Stationary and
Coflowing Receiving Fluids Journal of Hydraulic Engineering ASCE,
134(4), 390-402.

[21] Cuthbertson, A.J.S., Apsley, D.D., Davies, P.A., Lipari, G., Stansby, P.K.
(2008) Deposition from particle-laden, plane, turbulent, buoyant jets.
Journal of Hydraulic Engineering ASCE, 134(8), 1110-1122.

[22] Dantec Dynamics (2000). FlowManager software and Introduction to PIV


Instrumentation: Software Users guide.

[23] Doroodchi, E., Evans, G., Schwarz, M., Lane, G., Shah, N. and Nguyen,
A. (2008) Influence of turbulence intensity on particle drag coefficients.
Chemical Engineering Journal, 135, 129-134.

[24] EPD (2006). 20 Years of Marine Water Quality Monitoring in Hong Kong,
1986-2005. Environmental Protection Department, HKSAR.

[25] Ernst, G. G. J., Sparks, R. S. J., Carey, S. N., and Bursik, M. I. (1996).
Sedimentation from turbulent jets and plumes. Journal of Geophysical
Research B: Solid Earth, 101(3), 55755589.

202
[26] European Space Agency (2009). Volcanoes. In Space and Earth Monitor-
ing, retrieved 20 Jan 2013, from http://www.esa.int/.
[27] Farley, K.J. (1990). Predicting organic accumulation in sediments near
marine outfalls. Journal of Environmental Engineering, ASCE, 116(1), 144-
165.
[28] Fox, D.G. (1970). Forced plume in a stratified fluid. Journal of Geophysical
Research, 75, 6818-6835.
[29] Fischer, H.B., List, E.J., Koh, R.C.Y. Imberger, J. and Brooks, N.H. (1979).
Mixing in Inland and Coastal Waters. Academic Press.
[30] Gensheimer, R.J., Adams, E.E. and Law, A.W.K. (2012). Dynam-
ics of Particle Clouds in Ambient Currents with Application to Open-
Water Sediment Disposal. Journal of Hydraulic Engineering ASCE. doi:
10.1061/(ASCE)HY.1943-7900.0000659 (in press).
[31] Gomes, R.L. and Lester, J.N. (2003). Endocrine disrupters in receiving
waters. In: Endocrine Disrupters in Wastewater and Sludge Treatment
Processes, eds. J.W. Birkett and J.N. Lester, Lewis Publishers, 177-217.
[32] Gosman, A.D. and Ioannides, E. (1983). Aspects of computer simulation
of liquid-fueled combustors. Journal of Energy, AIAA, 7(6), 482-490.
[33] Hallermeier, R.J. (1981). Terminal settling velocity of commonly occurring
sand grains. Sedimentology, 28(6), 859-865.
[34] Ho, H. W. (1964). Fall velocity of a sphere in a field of oscillating fluid. PhD
thesis, State University of Iowa.
[35] Hwang, P.A. (1985). Fall velocity of particles in oscillating flow. Journal
of Hydraulic Engineering, ASCE, 111(3), 485-502.
[36] Ji, Z.G. (2008). Hydrodynamics and Water Quality: Modelling Rivers, Lakes
and Estuaries. Wiley-Interscience.
[37] Jiang, J., Law, A. W. K., and Cheng, N. S. (2005). Two-phase analysis
of vertical sediment-laden jets. Journal of Engineering Mechanics, 131(3),
308318.
[38] Kato, Y., Fujinaga, K., Nakamura, K., Takaya, Y., Kitamura, K., Ohta,
J., Toda, R., Nakashima, T. and Iwamori, H. (2011) Deep-sea mud in
the Pacific Ocean as a potential resource for rare-earth elements. Nature
Geoscience, 4(8), 535-539.
[39] Koh, R.C.Y. and Chang, Y.C. (1973) Mathematical Model for Barged Ocean
Disposal of Wastes, USEPA.
[40] Kuang, C.-P. and Lee, J.H.W. (2006) Stability and Mixing of a Vertical
Axisymmetric Buoyant Jet in Shallow Water. Environmental Fluid Me-
chanics, Vol.6, 153-180.

203
[41] Lai, A. C. H. (2009). Mixing of a Rosette Buoyant Jet Group. Ph.D. thesis,
The University of Hong Kong.

[42] Lai, A. C. H. and Lee, J. H. W. (2012a). Dynamic interaction of multiple


buoyant jets. Journal of Fluid Mechanics, 708, 539-575.

[43] Lai, C. C. K. and Lee, J. H. W. (2012b). Mixing of inclined dense jets in


stationary ambient. Journal of Hydro-environment Research, 6(1), 9-28.

[44] Lamb, H. (1932). Hydrodynamics. Cambridge University Press.

[45] Lane-Serff, G. F. and Moran, T. J. (2005). Sedimentation from buoyant


jets. Journal of Hydraulic Engineering, 131(3), 166174.

[46] Launder, B. E. and Spalding, D. B. (1974). The numerical computation of


turbulent flows. Computer Methods in Applied Mechanics and Engineering.
485, 3(2), 269-289.

[47] Lee, J.H.W. (1981). An hourglass experiment. International Journal of


Mechanics and Engineering Education 9(1), 39-46.

[48] Lee, J.H.W. and Cheung, V. (1990). Generalized lagrangian model for
buoyant jets in current. Journal of Environmental Engineering, ASCE,
116(6), 10851106.

[49] Lee, J.H.W. and Chu, V.H. (2003). Turbulent Jets and Plumes: A La-
grangian Approach. Kluwer Academic Publishers.

[50] Lee, J. H. W. and Jirka, G.H. (1981). Vertical round buoyant jet in Shallow
Water. Journal of Hydraulic Engineering, ASCE, 107(HY12), 16511675.

[51] Lee, K.W.Y. (2010). Mixing of Sediment-Laden Jet. Ph.D. thesis, The Uni-
versity of Hong Kong.

[52] Lee, K.W.Y., Li, A.C.Y. and Lee, J.H.W. (2012) Structure of a horizontal
sediment-laden momentum jet. Journal of Hydraulic Engineering ASCE.
doi:10.1061/(ASCE)HY.1943-7900.0000662 (in press).

[53] Li, A.C.Y. (2006). Theoretical Modeling and Experimental Studies of


Particle-Laden Plumes from Wastewater Discharges. M.Phil. thesis, The
University of Hong Kong.

[54] Li, C.W. (1997) Convection of particle thermals. Journal of Hydraulic


Research, 35(3), 363-376.

[55] Liang, L., Michaelides, E.E. (1992). The magnitude of Basset forces in
unsteady multiphase flow computations., ASME Journal of Fluids Engi-
neering, 114, 417-419.

[56] Metcalf, Eddy, I., Tchobanoglous, G., Burton, F. L., and Stensel, H. D.
(1991). Wastewater engineering: treatment and reuse. McGraw-Hill series
in civil and environmental engineering. McGraw-Hill, Boston, 2nd edition.

204
[57] Maxey, M.R. and Riley, J.J. (1983). Equation of motion for a small rigid
sphere in a nonuniform flow., Physics of Fluids, 26, 883-889.

[58] McLarnon, C.P. and Davies, P.A. (1999) Sedimentation of particles from
a finite volume horizontal buoyant jet. Proceedings of the 28th IAHR
congress, Graz, Austria, 22-27 Aug, 1999

[59] Mei, R. (1991). Effect of turbulence on the particle settling velocity in the
nonlinear drag range., International Journal of Multiphase Flow, 20(2),
273-284.

[60] Mei, R., Adrian, R.J., Hanratty, T.J. (1991). Particle dispersion in isotropic
turbulence under Stokes drag and Basset force with gravitational settling.,
Journal of Fluid Mechanics, 225, 481-495.

[61] Michaelides, E.E. (1992). A novel way of computing the Basset term in
unsteady multiphase flow computations., Physics of Fluids A, 7, 1579-1582.

[62] Michaelides, E.E. (1997). Review - the transient equation of motion for
particles, bubbles and droplets, ASME Journal of Fluids Engineering, 119,
233-247.

[63] Millero, F.J. and Poisson, A. (1981). International one-atmosphere equa-


tion of state of seawater. Deep Sea Research Part A, Oceanographic Re-
search Papers, 28(6), 625-629

[64] Minh, T.B., Leung, H.W., Loi, I.H., Chan, W.H., So, M.K., Mao, J.Q.,
Choi, D., Lam, J.C.W., Zheng, G., Martin, M., Lee, J.H.W., Lam, P.K.S.,
Richardson, B.J. (2009). Antibiotics in the Hong Kong metropolitan area:
Ubiquitous distribution and fate in Victoria Harbour. Marine Pollution
Bulletin, 58(7), 1052-1062.

[65] Morton, B.R., Taylor, G.I., and Turner, J.S. (1956). Turbulent gravita-
tional convection from maintained and instantaneous sources. Proceedings
of Royal Society of London, A234, 1-23.

[66] Murray, S. P. (1970). Settling velocities and vertical diffusion of particles


in turbulent water. Journal of Geophysical Research, 75(9), 1647-1654.

[67] Neves, M.J. and Fernando, H.J.S. (1995) Sedimentation of particles from
jets discharged by ocean outfalls: a theoretical and laboratory study. Water
Science and Technology, 32(2), 133-139.

[68] Nielsen, P. (1984). On the motion of suspended sand particles. Journal of


Geophysical Research, 89(C1), 616-626.

[69] Nielsen, P. (1992). Coastal Bottom Boundary Layers and Sediment Trans-
port. Advanced Series on Ocean Engineering - Vol.4. World Scientific.

[70] Nielsen, P. (1993). Turbulence effects on the settling of suspended parti-


cles. Journal of Sedimentary Petrology, 63(5), 835-838.

205
[71] Papanicolaou, P. N. and List, E. J. (1988). Investigations of round vertical
turbulent buoyant jets. Journal of Fluid Mechanics, 195, 341391.

[72] Parthasarathy, R. N. and Faeth, G. M. (1987). Structure of particle-laden


turbulent water jets in still water. International Journal of Multiphase
Flow, 13(5), 699716.

[73] Pedersen, F.B. (1980). A Monograph on Turbulent Entrainment and Friction


in Two-Layer Stratified Flow. Institute of Hydrodynamics and Hydraulic
Engineering, Technical University of Denmark, Series Paper No. 25.

[74] Pope, S.B. (2000). Turbulent Flows. Cambridge University Press.

[75] Raffel, M., Willert, C. and Kompenhans, J. (1998) Particle Image Velocime-
try: A Practical Guide. Springer

[76] Rodi, W. (1984). Turbulence models and their application in hydraulics: a


state of the art review. (2nd rev. ed.). International Association for Hydraulic
Research.

[77] Sato, Y. and Yamamoto, K. (1987). Lagrangian measurement of fluid-


particle motion in an isotropic turbulent field. Journal of Fluid Mechanics,
175, 183-199.

[78] Shih, T. H., Liou, W. W., Shabbir, A., Yang, Z., and Zhu, J. (1995). A
new k- eddy viscosity model for high Reynolds number turbulent flows.
Computers and Fluids, 24(3), 227238.

[79] Singamsetti, S. (1966). Diffusion of sediment in a submerged jet. Journal


of Hydraulics Division, Proceedings of ASCE, 92(2), 153-168.

[80] Smithsonian Institution (1954). Smithsonian Physical Tables. Smithsonian


Institution Press, 9th rev. ed.

[81] Snyder, W.H. and Lumley, J.L. (1971). Some measurements of particle
velocity autocorrelation functions in a turbulent flow. Journal of Fluid
Mechanics, 48(1), 41-71.

[82] Soulsby, R. (1997). Dynamics of marine sands : a manual for practical


applications. Thomas Telford, London.

[83] Sparks, R. S. J., Carey, S. N., and Sigurdsson, H. (1991). Sedimentation


from gravity currents generated by turbulent plumes. Sedimentology, 38(5),
839856.

[84] Stommel, H. (1949). Trajectories of small bodies sinking slowly through


convection cells. Journal of Marine Research, 8(1), 24-29.

[85] Stout, J.E., Arya, S.P. and Genikhovich, E.L. (1995). The effect of nonlin-
ear drag on the motion and settling velocity of heavy particles. Journal of
the Atmospheric Sciences, 52(22), 3836-3848.

206
[86] Taylor, G.I, (1921). Diffusion by continuous movements. Proceedings of
the London Mathematical Society, 20, 196-211.

[87] Thomann, R.V. and Mueller, J.A. (1987). Principles of Surface Water Qual-
ity Modeling and Control, Harper & Row.

[88] Tooby, P.A., Wick, G.L. and Issac, J.D. (1977). The Motion of a Small
Sphere in a Rotating Velocity Field: A Possible Mechanism for Suspending
Particles in Turbulence. Journal of Geophysical Research, 82(15), 2096-
2100.

[89] Toorman, E.A. (2002). Modelling of turbulent flow with suspended cohesive
sediment. in: Fine Sediment Dynamics in the Marine Environment, eds.
J.C. Winterwerp and C. Kranenburg, 155-169.

[90] Wang, H. and Law, A. W.-K. (2002). Second-order integral model for a
round turbulent buoyant jet. Journal of Fluid Mechanics, 459, 397428.

[91] Wood, I.R., Bell, R.G., Wilkinson, D.L. (1993). Ocean Disposal of Wastew-
ater, World Scientific

[92] Wygnanski, I. and Fiedler, H. (1969). Some measurements in the self-


preserving jet. Journal of Fluid Mechanics 38(3), 577-612

[93] Winterwerp, J.C. and van Kessel, T. (2003). Siltation by sediment-induced


density currents. Ocean Dynamics, 53, 186-196.

[94] Xiao, Y., Lee, J.H.W., Tang, H.-W. and Yu, D.Y. (2006) Three-
Dimensional Computations of Multiple Tandem Jets in Crossflow. China
Ocean Engineering, 20(1), 99-112.

[95] Xu, X.-R. and Li, X.-Y. (2010). Sorption and desorption of antibiotic tetra-
cycline on marine sediments. Chemosphere, 78, 430-436.

[96] Zarrebini, M. and Cardoso, S. S. S. (2000). Patterns of Sedimentation from


Surface Currents Generated by Turbulent Plumes. AIChE Journal, 46(10),
1947-1956.

[97] Zhang, J.J. and Li, X.Y. (2006). Coagulation and sedimentation of partic-
ulate pollutants in marine waters after wastewater outfall discharge. Water
Science and Technology: Water Supply, 6(1), 79-86.

[98] Zhou, Q. and Cheng, N.-S. (2009). Experimental investigation of single


particle settling in turbulence generated by oscillating grid. Chemical En-
gineering Journal, 149, 289300.

207
208
Appendix A

Numerical Solution of the


Equation of Particle Motion

A.1 General numerical solution


An analytical solution of the governing equation of particle motion (Chapter 3,
Eq. 3.1) can only be obtained in simplified flow fields with linearized drag forces.
Numerical integration is sought to solve the equation for more general cases, such
as in turbulent flow field. Rearranging the force terms, the equation of particle
motion can be simplified to
!
dup duf
(s + CM ) = (1 s)g + (1 + CM ) (A.1)
dt dt f
r
3CD 9
|up uf | (up uf ) + B
4d d
where s is the specific gravity s = p /f ; B is the integral in the Basset force
term. Eq. A.1 and the particle position (Eq. 3.7) are integrated using a predictor-
corrector approach. At time-step i with the particle position xi , particle velocity
up,i and fluid velocity uf ,i , an initial guess of particle velocity ufp,i+1 = up,i is
applied to estimate a predicted particle position xfi+1 at the next time step i + 1
(predictor step, denoted by superscript f).

xfi+1 = xi + ufp,i+1 t (A.2)

With this, the right hand side of Eq. A.1, representing the net forces of the
particle can be estimated using condition at the mid-point between xi and xfi+1 ,
including the fluid velocity and drag coefficient. Then an updated estimation of
the particle velocity ucp,i+1 can be estimated using the corrected right hand side
of Eq. A.1 (corrector step, denoted by superscript c), i.e.

ucp,i+1 = up,i + Fi+1/2 t (A.3)

where Fi+1/2 represents the right hand side of Eq. A.1 (gravity, drag, fluid ac-
celeration, added mass and Basset force terms) at the mid-step. A corrected

209
particle position of xci+1 at the next time-step can be reestimated and thus the
net forces. The iteration repeats until the convergence criteria is met. Iteration
stops when the difference of particle velocity magnitude at successive iterations
is less than 0.1%. Typically less than 5 iterations are required for convergence.

A.2 Numerical implementation of the Basset


force term
Evaluating the Basset force term requires the integration of the particle relative
velocity gradients throughout its time history
Z t p d(u u )
3 2 f

d f B, where B = d d
2 0 t
It poses two challenges to the solution of the equation of particle motion. Firstly,
the integral has to be evaluated every time step. As t increases, the numerical in-
tegration becomes increasingly cumbersome and time-consuming with additional
storage for the relative accelerations. Secondly, the integrand is ill-behaved as
t and becomes a infinity (Fig. A.1). Usual mathematical treatment like
subtracting out the singularity (see e.g. Acton, 1970) is not suitable here as the
relative velocity gradient d(upduf ) is unknown at time t. Although the integrand
is infinite at = t, the Basset integral is still finite.
The integral has to be evaluated by another approach. First, the integral is
decomposed into a sum of M integrals with each integrated within a small equal
time step t, as t = M t.
Z t dur Z t dur Z 2t dur Z M t dur
d d = d d + d d +...+ d d (A.4)
0 t 0 t t t (M 1)t t
where ur = up uf is the particle relative velocity. Secondly, the change of
relative velocity is evaluated with central difference within the time step t,
dur ur,i+1 ur,i
=
dt t
and the acceleration is assumed constant within the time interval, thus can be
separated out from the integral. The integral involving the square root is evalu-
ated as
Z kt q 
d
= 2 t (k 1)t t kt (A.5)
(k1)t t
 
= 2 t M k + 1 M k

The Basset integral B can be evaluated as a sum of all integrals


Z dur
t 2 MX 1 h i
B= d d = (ur,n+1 ur,n )( M n M n 1)
0 t t n=0
(A.6)

210
During the actual computation, particle velocity at time t is known and the
velocity at time t + t is going to be predicted. The Basset integral is integrated
from = 0 to = t + 12 t, as dur /dt is evaluated at the middle of the time step
(Fig. A.1).
Z t+t/2 dur
d
B = q d (A.7)
0 t + t/2
Z t dur Z t+t/2 dur
d d
= q d + q d
0 t + t/2 t t + t/2
d(up uf )
= Bt0 + 2t
dt
The last remaining integral of the half time step involves the acceleration at the
current time step, which is not known a priori. An implicit approach is used by
solving it together with the other terms in the equation to ensure stability. With
the Basset term Eq. A.1 can be written as
 r  !
9 dup duf
s + CM + 2t = (1 s)g + (1 + CM ) (A.8)
d dt dt f
r !
3CD 9 duf
|ur | ur Bt0 2t
4d d dt

where Bt0 is the Basset integral between = 0 t. Here the term involving
dup /dt in Eq. A.7 has been moved to the left hand side of the equation and to
be solved using a predictor-corrector approach using Eqs. A.2 and A.3. Finally
the accelerations are stored in arrays which are used for next time step compu-
tation. The validation of the Basset force formulation on the accelerated motion
of a spherical particle in a stagnant ambient is shown in Fig. 3.7 and shows
very good comparison with the analytical solution. Fig. A.2(a) shows the time
history of particle velocity for a spherical particle in a stagnant ambient with
its change, numerically solved with the Basset force term. The particle velocity
increases rapidly for the first 0.002s and the rate of increase slows down after-
wards. Fig. A.2(b) shows the Basset integrand which has an initial large value
but slowly decreases due to the reduction in particle acceleration. As t, the
integrand approaches rapidly to infinity.
Since the Basset force term is shown to be negligible for all sediment-laden jet
applications in this study, Eq. A.1 or Eq. A.8 can be simplified to the following
form excluding the Basset force term.
!
dup duf
(s + CM ) = (1 s)g + (1 + CM ) (A.9)
dt dt f
3CD
|up uf | (up uf )
4d

211
Figure A.1: Numerical integration of the Basset force term. Velocities at and
before time t is known and the velocity at time t + t is to be predicted. The
Basset integral is to be evaluated from 0 to t + 21 t.

212
1 1600
-1
wp ' = wp /ws dwp '/dt (s )
1400
0.8
wp' 1200

0.6 1000
800
0.4 600
400
0.2
dwp '/dt 200
t (s)
0 0
0 0.002 0.004 0.006 0.008 0.01

(a) Particle velocity and acceleration

(dwp'/dt) /
t-) (s-3/2)
1.0E+05
t = 0.002s
1.0E+04
t = 0.005s
1.0E+03

1.0E+02 t = 0.01s

1.0E+01
t (s)
1.0E+00
0 0.002 0.004 0.006 0.008 0.01

(b) Integrand of Basset integral

Figure A.2: Numerical solution of a particle falling in stagnant ambient, with


the Basset force term. d = 50m, p = 2500 kg/m3 , f = 1000 kg/m3 , ws =
2.0437mm/s, t = 104 s. (a) Time history of particle velocity and acceleration;
(b) the Basset integrand at t = 0.002, 0.005, 0.01s. The particle velocity wp is
normalized by the stillwater settling velocity ws .

213
214
Appendix B

3D Computation Fluid Dynamics


Modelling

B.1 Introduction
In this appendix, the details of the 3D computational fluid dynamics (CFD)
model for sediment-laden jet prediction are described. The CFD modelling serves
the following purposes:

1. Sediment concentration and deposition is predicted using the CFD model


and compared with experimental measurement and particle tracking model.
CFD modelling provides insights to understand the particle-turbulence in-
teraction which results in the reduction of particle settling velocity.
q
2. The CFD predicted RMS turbulent velocity fluctuations ( = 23 k, k =
turbulent kinetic energy) and the dissipation rate are used in the particle
tracking model for determining the turbulence length and time scales.

3. To provide numerical results for validating the integral model for spreading
current (Chapter 6).

B.2 Governing equations


The 3D Reynolds-averaged Navier-Stokes equations for incompressible steady
state are solved numerically. The Boussinesq approximation (density difference
is retained in the terms with gravitational acceleration) is adopted in buoyant
jet flows by virtue of small density difference. This governing equations are:
Continuity:
Ui
=0 (B.1)
xi
Momentum:
Ui 1 p ij f 0
Uj = + + gi (B.2)
xj f xi xj 0

215
where Ui is the fluid velocity in x, y and z directions; f is the density of jet
fluid; 0 is the ambient density; p is the dynamic pressure; g is the gravitational
acceleration; ij is the Reynolds stresses estimated as with the gradient transport
analogy: !
Uj Ui 2
ij = t + kij (B.3)
xi xj 3
where k is the turbulent kinetic energy; is the dissipation rate of k; t is the
turbulent eddy viscosity. A linear relationship between temperature and water
density is prescribed for the modelling of density-driven buoyant jets.
The equation for sediment transport, assuming very dilute concentration, is:
!
C t C
(Ui ws ) = (B.4)
xi xi Sct xi
where C is the sediment concentration and Sct is the turbulent Schmidt number
taken as 0.85, as default in the code. The term ws C
z
models the settling flux of
sediment, where ws is the settling velocity.

B.3 Turbulence closure


The time averaging process of the Navier-Stokes equations results in additional
terms as the cross-correlation of the velocity fluctuations (the turbulent shear
stresses, e.g, u v ). The turbulent stresses arising from the fluctuating velocities
can be modelled as the product of the mean velocity gradient and the turbulent
viscosity (see e.g. Rodi, 1984), i.e.
u
= u v = t
y
Two-equation k turbulence closure model is used for estimating the turbulent
eddy viscosity t . In the standard k model (Launder and Spalding, 1974),
the equations for turbulent kinetic energy k and the dissipation rates are
!
k t k
Ui = + Gk + Gb (B.5)
xi xi k xi
!
t 2
Ui = + C1 (Gk + (1 C3 )Gb ) C2 (B.6)
xi xi xi k k
where Gk and Gb are the production of k due to shear and buoyancy respectively:
!
Ui Uj Uj
Gk = t + (B.7)
xj xi xi
t
Gb = gi (B.8)
h xi
The turbulent eddy viscosity t is estimated as
k2
t = C (B.9)

216
The constants C , C1 , C2 , C3 , k , and h are determined by fitting the model
with experimental data of basic shear flows. In the standard k model, the
constants are C = 0.09, C1 = 1.44, C2 = 1.92, C3 = 1.0 for stable stratification
and zero for unstable stratification, k = 1.0, = 1.3 and h = 0.7. It is
reported (Rodi, 1984; Kuang and Lee, 2006) that the standard k model
cannot satisfactorily predict the spreading rate of axisymmetric jets.
In the present study, the turbulent closure utilizes the realizable k turbu-
lence closure model (Shih, 1995). The model differs from the standard k model
in two important ways: i) The eddy viscosity coefficient C is not a constant,
but determined as a function of k, , and the local strain rate;
1
C = (B.10)
A0 + As kU

where q
U = Sij Sij (B.11)
and
A0 = 4.04, As = 6 cos (B.12)
1 Sij Sjk Ski e q Ui Uj
= cos1 ( 6W ), W = ,S = Sij Sij , Sij = +
3 Se3 xj xi
and ii) a new transport equation for the dissipation rate is used
!
t 2
Ui = + C1 S C2 + C1 C3 Gb (B.13)
xi xi xi k + k

where !
k q
C1 = max 0.43, , = S , S = 2Sij Sij
+5
and C2 = 1.9, C1 = 1.44 and C3 = 1.0 for stable stratification and zero for
unstable stratification are the empirical constants. The realizable k is shown
to give more accurate prediction of the spreading rate for both planar and round
jets (Shih, 1995). The model has been successfully applied in the study of mul-
tiple tandem jets in crossflows (Xiao et al. , 2006). For the present jet flow
cases, the predicted spreading rate of jet half width is 0.12, comparable with the
experimentally measured value of 0.114 and better than the value of 0.13-0.14
obtained using standard k model (Fig. B.5).

B.4 Model of horizontal sediment-laden jet


B.4.1 Grid configuration
A three-dimensional unstructured mesh with 182,420 non-uniform hexahedral
elements is used to discretise the half-experimental tank and jet nozzle with
diameter of 6mm (Fig. B.1). The smallest mesh size in the direction of the jet is
6mm, about 1/10 of the length of the zone of flow establishment. The smallest

217
transverse mesh size at the jet nozzle is of size 0.6mm, 1/10 the diameter of the
jet nozzle. The jet tube is not separately modelled but located on the same plane
with the wall of the jet inlet.

Figure B.1: Computation mesh for horizontal sediment-laden jets

B.4.2 Boundary and initial conditions


Inflow velocity and sediment concentration are prescribed at the jet nozzle. The
turbulent kinetic energy k is estimated from well-established relations for fully
developed smooth pipe flow, k = 23 (uI)2 , where u = inflow velocity, I = tur-
bulence intensity related to the Reynolds number via the Blassius equation
3/2
I = 0.16Re1/8 . The value of at the inlet is then given by = C3/4 k l ; l
is the length scale of energy containing eddies and adopted as l = 0.07D similar
to mixing length. For the tank bottom and tank wall adjacent to the jet nozzle,
a no slip wall boundary is used. For other boundaries, a zero pressure condition
is used, and k and are specified as 1% of the values of the jet inlet. By virtue
of symmetry about the vertical jet centerline plane (the x z plane), only half of
the tank is modeled. For the sediment transport, a zero flux boundary condition
is applied on all boundaries except at the bottom where a sediment flux ws C is
prescribed.
Zero velocity, pressure and sediment concentration are prescribed as initial
values for iteration. The initial k and in the whole domain are specified as 0.01
times the values of the inlet. Sensitivity tests show that the prescribed initial

218
values of k and do not have a significant effect on the final steady flow, since
the turbulence in the tank will evolve as the flow develops from the inlet.

B.4.3 Computational details


The governing equations are solved numerically using the finite volume method
as embodied in the FLUENT software. Velocities and other variables are defined
at the cell center. Cell face dependent variables are interpolated from the cell
center values using a second order upwind scheme. The SIMPLEC algorithm is
employed for velocity-pressure correction with under-relaxation factor of 0.3 for
pressure, 0.7 for momentum, 0.8 for k and , and 1.0 for turbulent viscosity and
sediment concentration. Convergence is declared when the normalized residual
is less than 106 for sediment concentration and 103 for all other variables.
Typically about 5000 iterations are required for convergence.

B.4.4 Model results


The model predictions on sediment-laden jets have been presented in Chapters
5 and 6. Here the single-phase jet results are presented.

Horizontal pure jet

Fig. B.2 shows the CFD predicted velocity and tracer concentration field at the
centerline vertical plane, showing the decay of jet velocity and tracer concentra-
tion. Fig. B.3 shows the predicted cross-sectional tracer concentration profiles
which reveals the self-similarity of concentration profiles in different jet cross-
sections. Fig. B.4 shows the comparison between predicted and theoretical jet
centerline velocity. Fig. B.5 shows the comparison between CFD predicted Gaus-
sian half-width of a pure round jet and the commonly adopted spreading rate of
0.114. The prediction using the realizable k model compares well with the
theoretical value.

Horizontal buoyant jet

Fig. B.6 shows the computed centerline vertical plane velocity of horizontal buoy-
ant jets of two cases of Li (2006)s experiments. Fig. B.7 shows the predicted jet
trajectories using CFD and integral model JETLAG and both model predictions
compares well. The predicted dilution also compares with the experimentally
derived Cederwall equation (Cederwall, 1968) (Fig. B.8).
z z
Sc = 0.54F r(0.38 + 0.66)5/3 for > 0.5 (B.14)
F rD F rD
z 7/16 z
Sc = 0.54F r( ) for 0.5 (B.15)
F rD F rD
where Sc is the minimum dilution defined at jet centerline, F r is the initial jet
densimetric Froude number and D is the jet diameter.

219
0.04
0.3 m/s

0.02
(m)

-0.02

-0.04 (a) Velocity


0 0.1 0.2 0.3
(m)

0.04

0.05
0.02
0.1 0.1
0.05
0.15 0.5 0.3 0.2 0.15
(m)

0
0.15 0.3 0.6 0.4
0.2
00.1.0
5 0.15
-0.02 0.1
0.05
-0.04
(b) Tracer
Concentration
0 0.1 0.2 0.3
(m)

Figure B.2: CFD predicted velocity and tracer concentration (C/C0 ) field of
x z plane at y = 0. u0 = 0.41 m/s; D = 6 mm; C0 = 5.29 g/L. The dashed
lines are the jet top-hat boundary defined as bT = 0.16x.

0.04 0.04
x = 10D = 0.06m x = 30D = 0.18m
0.1
0.03
1 2 0.2
0. 0 .
0.02 0.02
0.4 0.4
0.1

6 0.
0.01 0. 6
0.4
0.1

0.
(m)

(m)

0.2
8

0 0
0.8

0.4
0.2
0.4

0.1

0.8

0.4
-0.01 0.2 .1
0
0.6
0.6
0.1

-0.02 -0.02
0.4
0.2 0.4
-0.03 0.1 0.2
0.1
-0.04 -0.04
-0.04 -0.02 0 0.02 0.04 -0.04 -0.02 0 0.02 0.04
(m) (m)

Figure B.3: CFD predicted tracer concentration field (C/Cc ) of y z plane at


x = 10D and x = 30D. u0 = 0.41 m/s; D = 6 mm. The dashed circles are the
jet top-hat boundary defined as bT = 0.16x.

220
uc /u0
10
Theoretical
FJ58, u0 = 0.57 m/s
FJ42, u0 = 0.41 m/s

x/D
0.1
1 10 100

Figure B.4: Comparison of CFD predicted and theoretical centerline velocity


uc /u0 = 6.2(x/D)1 .

7
FJ42
6 FJ58
b = 0.114x
5

4
b/D

0
0 10 20 30 40 50 60
x/D

Figure B.5: Predicted jet Gaussian half width using the realizable k model.
= 0.114 is the commonly adopted value for round jets.

221
0.4
(i) x = 0 - 0.2m (ii) x = 0.15 - 0.55m
0.5 m/s 0.1 m/s

0.35
0.1

0.3

0.25
0.05
(m)

(m)
0.2

0.15
0

0.1

0.05
-0.05
0 0.05 0.1 0.15 0.2 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55
(m) (m)

(a) CB66, u0 =0.65 m/s, F r = 19.5

0.5 m/s 0.35 0.1 m/s

0.1
0.3

0.25

0.05
(m)

(m)

0.2

0.15

0
0.1

0.05
(i) x = 0 - 0.2m (ii) x = 0.1 - 0.45m
-0.05
0 0.05 0.1 0.15 0.2 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
(m) (m)

(b) CB50, u0 =0.49 m/s, F r = 14.9

Figure B.6: CFD predicted horizontal buoyant jet velocities. (i) Initial region,
(ii) bend up region. The dashed lines are the JETLAG predicted jet boundary.

222
0.4
CB50, Fr=14.9
0.35 CB54, Fr=16.4
CB66, Fr=19.5
0.3
JetLag
0.25
z (m)

0.2

0.15

0.1

0.05

0
0 0.1 0.2 0.3 0.4 0.5
x (m)

Figure B.7: Comparison of computed horizontal buoyant jet trajectories of CFD


and JETLAG.

10
CB50, Fr=14.9
CB54, Fr=16.4
CB66, Fr=19.5
Cederwall
Sc/Fr

0.1
0.01 0.1 1 10
z/FrD

Figure B.8: Comparison of CFD computed horizontal buoyant jet centerline


dilutions with Cederwall equation.

223
B.5 Model for spreading current
The three-dimensional flow of a vertical axisymmetric buoyant jet impinging on
a free surface in stagnant water is obtained by solving the Reynolds-averaged
Navier-Stokes equations (RANS) assuming axisymmetry. The model domain
is 0.53m (r) 0.3m (z). The model grid consists of 10000 2D cells with both
100 cells in the radial and axial direction. The half-jet inlet (D = 7mm) is
correspondingly resolved by 5 cells. Grid resolution increases towards the jet
axis and the surface layer. The free surface is modelled by symmetric boundary
condition to provide zero velocity gradient and zero mass flux condition. The
bottom boundary is modelled as a no-slip wall and the remaining side is modelled
as an open boundary of zero excess pressure (Fig. B.9).
Figs. B.10 and B.11 show the computed flow field and tracer concentration
field at the spreading layer region for two cases, illustrating the transition from
the vertical jet to the radial spreading current. There is a weak internal hydraulic
jump in the spreading layer as depicted by the slight increase in thickness of the
current. The tracer concentration decays in the spreading layer slowly due to
the much weaker entrainment by the spreading current. The comparison of CFD
and integral model predictions on spreading layer velocity and reduced gravity
can be found in Chapter 6.

Free surface (symmetric)


0.3

0.2

Axis Open
z (m)

0.1

0
Wall

Jet

-0.1
0 0.1 0.2 0.3 0.4 0.5
r (m)

Figure B.9: Computation mesh for vertical axisymmetric buoyant jet impinging
the surface.

224
(a) velocity
0.3

0.25
z (m)

0.2

0.1 m/s

0 0.05 0.1 0.15 0.2 0.25 0.3


r (m)

(b) concentration
0.3 0.05
0.025
0.01
0.075

0.025
0.

0.05
1

0.01

0.25
z (m)

0.075
0.1

0.025
0.05

0.2
0.01
0.125 .15
0

0 0.05 0.1 0.15 0.2 0.25 0.3


r (m)

Figure B.10: CFD predicted spreading layer velocity field and concentration field
(C/C0 ). u0 = 0.351m/s, D = 7mm, F r = 10.0.

225
(a) velocity
0.3

0.25
z (m)

0.2

0.1 m/s

0 0.05 0.1 0.15 0.2 0.25 0.3


r (m)

(b) concentration
0.3 5
0.0
0.01
5
0.02
0.01

0.25
0.05
z (m)

0.025

0.2
0.05

0.01
0.075
0.1

0 0.05 0.1 0.15 0.2 0.25 0.3


r (m)

Figure B.11: CFD predicted spreading layer velocity field and concentration field
(C/C0 ). u0 = 0.146m/s, D = 7mm, F r = 4.0.

226
Appendix C

Particle Settling Velocity in


Ethanol-Water Mixture

In Chapter 6, the horizontal sediment-laden buoyant jet experiments of Li (2006)


and Lee (2010) used ethanol-water mixture as a buoyancy source for the jet
(Tables 6.4-6.6). Density of source fluid is measured using density meter. Source-
ambient density difference are about 20-40 kg/m3 , which indicates that ethanol
concentration is 10-30% by weight (Fig. C.1). It is well known that the fluid
viscosity in ethanol-water mixture in this concentration range can be up to 2-3
times that of pure/tap water (Fig. C.2). The particle settling velocity in ethanol-
water mixture (jet fluid) can be much less than its counterpart in the tap water
ambient. Two examples are used to illustrate the effect of ethanol-water mixture
on sediment settling velocity (Figs. C.3 and C.4). In these two cases, the source
fluid viscosity is about 2 times of ambient viscosity and gradually decreases due to
turbulent mixing. The estimated sediment settling velocity at the jet centerline
is about half of that in the ambient.
The local settling velocity for the buoyant jet experiments of Li (2006) and
Lee (2010) is estimated based on the local density and molecular viscosity in the
particle tracking model and CFD simulations. In the particle tracking model,
the local fluid density is predicted using JETLAG; the ethanol concentration (%
by weight) is then back-estimated using Fig. C.1. The molecular viscosity of the
fluid is then estimated using Fig. C.2. The graphical relationships of density
and dynamic viscosity versus the percentage by weight of ethanol in water and
temperature are adopted from Smithsonian Physical Tables (Smithsonian Insti-
tution, 1954). A sensitivity study shows that using the locally estimated settling
velocity, the model prediction compares better with measured data than using a
single settling velocity based on the ambient density and viscosity (Fig. C.5).
For experiments using saline water as the ambient and freshwater as the
buoyancy source (e.g. Sparks et al. 1991; Ernst et al. 1996), the difference in
viscosity between saline water and freshwater are less than 5% (at 20 C, =
1.004 106 m2 /s for freshwater; = 1.04 106 m2 /s for seawater, salinity =
33ppt, f = 1023 kg/m3 ); a constant viscosity of freshwater is used for computing
the settling velocity in these cases.

227
3
Density (g/cm )
1

0.99

0.98

0.97

Temperature C
o
0.96

0.95

0.94 15
20
0.93 25
30
% by weight of ethanol
0.92
0 5 10 15 20 25 30 35 40

Figure C.1: Density of ethanol-water mixture.

4
Dynamic viscosity o
Temperature C
-3
3.5 (10 kg/m/s) 15

3
20

2.5
25

2 30

1.5

0.5
% by weight of ethanol
0
0 5 10 15 20 25 30 35 40

Figure C.2: Dynamic viscosity of ethanol-water mixture.

228
3 -3
c (kg/m ) c (10 kg/m/s)
1010 2.0
a
1000
1.5
990

980 1.0
c c
970
0.5
960
x (m)
950 0.0
0 0.05 0.1 0.15 0.2 0.25 0.3
(a) Centerline jet fluid density and viscosity

ws (cm/s)
1.6
ws - ambient
1.4
1.2
1
0.8
0.6
0.4
0.2
x (m)
0
0 0.05 0.1 0.15 0.2 0.25 0.3
(b) Estimated sediment settling velocity at jet centerline.

Figure C.3: Variation of settling velocity in a horizontal buoyant jet using


ethanol-water mixture as the source of buoyancy. Case FB58 (Li, 2006):
u0 = 0.57m/s; F r = 14.5; d = 133m (sand). (a) JETLAG predicted cen-
terline jet fluid density (c ) and estimated dynamic viscosity (c ) using Figs. C.1
& C.2. (b) Estimated sediment settling velocity at jet centerline using Soulsby
equation.

229
3 -3
c (kg/m ) c (10 kg/m/s)
1010 2.5
a
1000 2.0
990
1.5
980 c
c
1.0
970

960 0.5
x (m)
950 0.0
0 0.05 0.1 0.15 0.2 0.25 0.3
(a) Centerline jet fluid density and viscosity

ws (cm/s)
2.5
ws - ambient
2

1.5

0.5
x (m)
0
0 0.05 0.1 0.15 0.2 0.25 0.3
(b) Estimated sediment settling velocity at jet centerline.

Figure C.4: Variation of settling velocity in a horizontal buoyant jet using


ethanol-water mixture as the source of buoyancy. Case G180B90Fr17 (Lee, 2010):
u0 = 0.88m/s; F r = 18.3; d = 180m (glass). (a) JETLAG predicted centerline
jet fluid density (c ) and estimated dynamic viscosity (c ) using Figs. C.1 &
C.2. (b) Estimated sediment settling velocity at jet centerline using spherical
drag law.

230
0.5
F (g/m/s) FB58
0.4 Local ws
Ambient ws
0.3
Meas.
0.2

0.1

x (m)
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
(a) Case FB58 (Li, 2006): u0 = 0.57m/s; F r = 14.5; d = 133m .

0.5
Fs (g/m/s) G180B90Fr17
0.4 Local ws
Ambient ws
0.3
Meas.
0.2

0.1
x (m)
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
(b) Case G180B90Fr17 (Lee, 2010): u0 = 0.88m/s; F r = 18.3; d =
180m .

Figure C.5: Model comparison of using local settling velocity and ambient set-
tling velocity in the prediction of sediment deposition on buoyant jets.

231

You might also like