0% found this document useful (0 votes)
227 views318 pages

OE4607 Introduction Dredging Engineering 2 ND Edition

Uploaded by

Sandeep Rao
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
227 views318 pages

OE4607 Introduction Dredging Engineering 2 ND Edition

Uploaded by

Sandeep Rao
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 318

OE4607 Introduction Dredging Engineering.

MSc Offshore & Dredging Engineering


Delft University of Technology
2nd Edition

by
Dr.ir. Sape A. Miedema

Copyright © Dr.ir. S.A. Miedema


Introduction Dredging Engineering.

Preface
In dredging, trenching, (deep sea) mining, drilling, tunnel boring and many other applications, sand, clay or rock
has to be excavated. The productions (and thus the dimensions) of the excavating equipment range from mm 3/sec
- cm3/sec to m3/sec. In oil drilling layers with a thickness of a magnitude of 0.2 mm are cut, while in dredging this
can be of a magnitude of 0.1 m with cutter suction dredges and meters for clamshells and backhoe’s. Some
equipment is designed for dry soil, while others operate under water saturated conditions. Installed cutting powers
may range up to 10 MW. For both the design, the operation and production estimation of the excavating equipment
it is important to be able to predict the cutting forces and powers. After the soil has been excavated it is usually
transported hydraulically as a slurry over a short (TSHD’s) or a long distance (CSD’s) or mechanically. Estimating
the pressure losses and determining whether or not a bed will occur in the pipeline is of great importance.
Fundamental processes of sedimentation, initiation of motion and erosion of the soil particles determine the
transport process and the flow regimes. In TSHD’s the soil has to settle during the loading process, where also
sedimentation and erosion will be in equilibrium. In all cases we have to deal with soil and high density soil water
mixtures and its fundamental behavior.

The lecture notes are complemented with the book The Delft Sand, Clay & Rock Cutting Model, by Dr.ir. Sape
A. Miedema and available at www.iospress.nl. The first version is used at the exam. The newest edition is available
on ResearchGate.
The book Slurry Transport: Fundamentals, A Historical Overview & The Delft Head Loss & Limit Deposit
Velocity Framework is also available on ResearchGate.

Additional information can be found on www.delftdredging.com, www.dredgingengineering.com,


www.dhlldv.com and www.dscrcm.com.

Copyright © Dr.ir. S.A. Miedema


Introduction Dredging Engineering.

Targets/Goals of OE4607.
1. What is dredging (projects and equipment)?
2. What are the important dredging processes?
3. The working methods of a CSD and a TSHD.
4. Which dredging processes can be identified related to a CSD and TSHD?
a. CSD.
b. TSHD.
5. Basic Soil Mechanics (OE4624-15 for advanced theory).
a. Which soil mechanical properties are involved?
b. Failure criteria.
c. Soil mechanical tests.
d. Mohr circles, active and passive soil failure.
6. Basic saturated sand cutting theory (OE4626 for advanced theory and OE4627 for jetting).
a. The forces involved.
b. The generic force and moment equilibrium equations.
c. Dilatation in saturated sand.
d. Simplified cutting equations.
e. Simplified specific energy equations.
f. The terminal settling velocity (including hindered settling).
7. The terminal setting velocity.
a. The force equilibrium and theoretical equation.
b. The drag coefficient.
c. Practical equations.
8. Volume, volume flow, concentration and density relations.
9. Basic pumps and slurry transport (OE4625 for advanced theory).
a. Flow (Q)-Head (H) curves.
b. Darcy Weisbach and ELM for pure liquid.
c. The Dacry Weisbach friction factor.
d. The DHLLDV Framework, Jufin Lopatin and Wilson for mixture.
e. The Limit Deposit Velocity.
10. Basic hopper sedimentation and overflow losses (OE4627 for advanced theory).
a. The TSHD loading cycle.
b. The optimum loading time.
c. The Camp theory.
d. The settling efficiency.
e. Overflow losses.
11. Case study CSD, showing the relation between the different processes.
12. Case study TSHD, showing the relation between the different processes.

Book ISBN: 978-94-6186-536-6


EBook ISBN: 978-94-61864-57-4
DOI: 10.13140/RG.2.1.2643.2488

Version: Tuesday, August 16, 2016

LEGAL NOTICE
The publisher and/or author are not responsible for the use which might be made of the following information.

PRINTED IN THE NETHERLANDS

Copyright © Dr.ir. S.A. Miedema


Introduction Dredging Engineering.

About the Author.


Dr.ir. Sape A. Miedema (November 8th 1955) obtained his
M.Sc. degree in Mechanical Engineering with honours at the
Delft University of Technology (DUT) in 1983. He obtained
his Ph.D. degree on research into the basics of soil cutting in
relation with ship motions, in 1987. From 1987 to 1992 he was
Assistant Professor at the chair of Dredging Technology. In
1992 and 1993 he was a member of the management board of
Mechanical Engineering & Marine Technology of the DUT. In
1992 he became Associate Professor at the DUT with the chair
of Dredging Technology. From 1996 to 2001 he was appointed
Head of Studies of Mechanical Engineering and Marine
Technology at the DUT, but still remaining Associate Professor of Dredging Engineering. In
2005 he was appointed Head of Studies of the MSc program of Offshore & Dredging
Engineering and he is also still Associate Professor of Dredging Engineering. In 2013 he was
also appointed as Head of Studies of the MSc program Marine Technology of the DUT.

Dr.ir. S.A. Miedema teaches (or has taught) courses on soil mechanics and soil cutting, pumps
and slurry transport, hopper sedimentation and erosion, mechatronics, applied thermodynamics
related to energy, drive system design principles, mooring systems, hydromechanics and
mathematics. He is (or has been) also teaching at Hohai University, Changzhou, China, at
Cantho University, Cantho Vietnam, at Petrovietnam University, Baria, Vietnam and different
dredging companies in the Netherlands and the USA.
His research focuses on the mathematical modeling of dredging systems like, cutter suction
dredges, hopper dredges, clamshell dredges, backhoe dredges and trenchers. The fundamental
part of the research focuses on the cutting processes of sand, clay and rock, sedimentation
processes in Trailing Suction Hopper Dredges and the associated erosion processes. Lately the
research focuses on hyperbaric rock cutting in relation with deep sea mining and on hydraulic
transport of solids/liquid settling slurries.

Copyright © Dr.ir. S.A. Miedema


Introduction Dredging Engineering.

Copyright © Dr.ir. S.A. Miedema


Introduction Dredging Engineering.

Table of Contents
Chapter 1: Basic Soil Mechanics. .................................................................................................................. 1
1.1. Introduction. .......................................................................................................................................... 1
1.2. Soil Mechanics. ..................................................................................................................................... 1
1.2.1. Definition. ......................................................................................................................................... 1
1.2.2. Soil Creation. ..................................................................................................................................... 3
1.2.3. Soil Classification. ............................................................................................................................ 4
1.3. Soils. ...................................................................................................................................................... 7
1.3.1. Sand. .................................................................................................................................................. 7
1.3.2. Clay. .................................................................................................................................................. 9
1.3.3. Rock. ................................................................................................................................................11
1.4. Soil Mechanical Parameters. ................................................................................................................17
1.4.1. Grain Size Distribution/Particle Size Distribution. ..........................................................................17
1.4.2. Atterberg Limits. ..............................................................................................................................17
1.4.2.1. Shrinkage Limit. ......................................................................................................................18
1.4.2.2. Plastic Limit. ...........................................................................................................................18
1.4.2.3. Liquid Limit. ...........................................................................................................................18
1.4.2.4. Importance of Liquid Limit Test. ............................................................................................19
1.4.2.5. Derived Limits. ........................................................................................................................19
1.4.2.6. Plasticity Index. .......................................................................................................................19
1.4.2.7. Liquidity Index. .......................................................................................................................19
1.4.2.8. Activity. ...................................................................................................................................19
1.4.3. Mass Volume Relations. ..................................................................................................................20
1.4.3.1. Specific Gravity. ......................................................................................................................20
1.4.3.2. Density.....................................................................................................................................20
1.4.3.3. Relative Density. .....................................................................................................................21
1.4.3.4. Porosity. ...................................................................................................................................23
1.4.3.5. Void ratio. ................................................................................................................................23
1.4.3.6. Dilatation. ................................................................................................................................23
1.4.4. Permeability. ....................................................................................................................................24
1.4.5. The Angle of Internal Friction..........................................................................................................26
1.4.6. The Angle of External Friction. .......................................................................................................27
1.4.7. Shear Strength. .................................................................................................................................28
1.4.7.1. Introduction. ............................................................................................................................28
1.4.7.2. Undrained Shear Strength. .......................................................................................................28
1.4.7.3. Drained Shear Strength. ...........................................................................................................28
1.4.7.4. Cohesion (Internal Shear Strength). ........................................................................................29
1.4.7.5. Adhesion (External Shear Strength). .......................................................................................29

Copyright © Dr.ir. S.A. Miedema


Introduction Dredging Engineering.

1.4.8. UCS or Unconfined Compressive Strength. .....................................................................................30


1.4.9. Unconfined Tensile Strength. ...........................................................................................................31
1.4.10. BTS or Brazilian Tensile Strength. ..............................................................................................31
1.4.11. Hardness. .....................................................................................................................................31
1.5. Criteria & Concepts. .............................................................................................................................33
1.5.1. Failure Criteria. ................................................................................................................................33
1.5.2. The Phi=0 Concept. ..........................................................................................................................33
1.5.3. Factors Controlling Shear Strength of Soils. ....................................................................................33
1.5.4. Friction, Interlocking & Dilation. .....................................................................................................34
1.5.5. Effective Stress. ................................................................................................................................34
1.5.6. Pore Water Pressure: Hydrostatic Conditions. .................................................................................35
1.5.7. Pore Water Pressure: Capillary Action. ............................................................................................35
1.5.8. Darcy’s Law. ....................................................................................................................................35
1.5.9. Brittle versus Ductile Failure. ..........................................................................................................37
1.6. Soil Mechanical Tests. ..........................................................................................................................39
1.6.1. Sieve Analysis. .................................................................................................................................39
1.6.2. Hydrometer Analysis. .......................................................................................................................39
1.6.3. Standard Penetration Test. ................................................................................................................40
1.6.4. Cone Penetration Test. .....................................................................................................................41
1.6.5. Triaxial Test. ....................................................................................................................................43
1.6.5.1. Consolidated Drained (CD). ....................................................................................................45
1.6.5.2. Consolidated Undrained (CU). ................................................................................................45
1.6.5.3. Unconsolidated Undrained (UU). ............................................................................................45
1.6.6. Shear Test. ........................................................................................................................................45
1.6.7. Point Load Test. ...............................................................................................................................46
1.7. Nomenclature. ......................................................................................................................................49
1.8. Notes. ....................................................................................................................................................51
Chapter 2: The Terminal Settling Velocity of Particles. ............................................................................57
2.1. Introduction. .........................................................................................................................................57
2.2. The Equilibrium of Forces. ...................................................................................................................57
2.3. The Drag Coefficient. ...........................................................................................................................58
2.4. Terminal Settling Velocity Equations...................................................................................................61
2.5. The Shape Factor ..................................................................................................................................65
2.6. Hindered Settling. .................................................................................................................................67
2.7. Conclusions. .........................................................................................................................................69
2.8. Nomenclature. ......................................................................................................................................70
2.9. Notes. ....................................................................................................................................................71
Chapter 3: The Pump/Pipeline System. .......................................................................................................77
3.1. Introduction. .........................................................................................................................................77
3.2. The Pump Drive....................................................................................................................................78
3.3. The Centrifugal Pump. .........................................................................................................................79
3.4. Affinity Laws. .......................................................................................................................................81

Copyright © Dr.ir. S.A. Miedema


Introduction Dredging Engineering.

3.5. Approximations. ...................................................................................................................................83


3.6. The Total Head Losses. ........................................................................................................................83
3.7. The Pump/Pipeline System Description. ..............................................................................................85
3.8. The Segmented Pipeline. ......................................................................................................................85
3.9. The Inertial Effects in the Pipeline. ......................................................................................................86
3.10. Case study. ............................................................................................................................................89
3.11. Conclusions and Discussion. ................................................................................................................96
3.12. Nomenclature. ......................................................................................................................................97
3.13. Notes. ....................................................................................................................................................99
Chapter 4: Pressure Losses with Homogeneous Liquid Flow. ................................................................105
4.1. Pipe Wall Shear Stress. .......................................................................................................................105
4.2. The Darcy-Weisbach Friction Factor. ................................................................................................106
4.3. The Equivalent Liquid Model. ............................................................................................................107
4.4. Approximation of the Darcy-Weisbach Friction Factor. ....................................................................108
4.5. The Friction Velocity or Shear Velocity u*. .......................................................................................108
4.6. The Thickness of the Viscous Sub Layer δv. ......................................................................................108
4.7. The Smallest Eddies. ..........................................................................................................................108
4.8. The Apparent Viscosity. .....................................................................................................................110
4.9. Nomenclature. ....................................................................................................................................114
4.10. Notes. ..................................................................................................................................................115
Chapter 5: The Delft Head Loss & Limit Deposit Velocity Framework. ...............................................121
5.1. Introduction. .......................................................................................................................................121
5.1.1. Considerations. ...............................................................................................................................121
5.1.2. Energy Dissipation. ........................................................................................................................124
5.1.3. Starting Points. ...............................................................................................................................128
5.1.4. Approach. .......................................................................................................................................132
5.1.5. Nomenclature Introduction.............................................................................................................134
5.2. Flow Regimes and Scenario’s. ...........................................................................................................135
5.2.1. Introduction. ...................................................................................................................................135
5.2.2. Concentration Considerations. .......................................................................................................136
5.2.3. The 8 Flow Regimes Identified. .....................................................................................................138
5.2.4. The 6 Scenario’s Identified. ...........................................................................................................142
5.2.4.1. Scenarios L1 & R1. ...............................................................................................................143
5.2.4.2. Scenarios L2 & R2. ...............................................................................................................145
5.2.4.3. Scenarios L3 & R3. ...............................................................................................................147
1.1.1.1 Conclusions & Discussion. ....................................................................................................149
5.2.5. Verification & Validation. ..............................................................................................................149
5.2.5.1. L1: Fixed Bed & Heterogeneous, Constant Cvs. ....................................................................150
5.2.5.2. R1: Heterogeneous, Constant Cvt...........................................................................................151
5.2.5.3. L2: Fixed & Sliding Bed – Heterogeneous & Sliding Flow, Constant Cvs. ...........................152
1.1.1.2 R2, R3: Sliding Bed & Sliding Flow, Constant Cvt. ..............................................................153
5.2.5.4. L1, R1, L2, R2:, Homogeneous. ............................................................................................154

Copyright © Dr.ir. S.A. Miedema


Introduction Dredging Engineering.

5.2.5.5. L3, R3: Sliding Bed & Sliding Flow, Constant C vs. ..............................................................155
5.2.6. Discussion & Conclusions..............................................................................................................156
5.2.7. Nomenclature Flow Regimes & Scenario’s. ..................................................................................157
5.3. Notes. ..................................................................................................................................................159
Chapter 6: Slurry Transport Models. .......................................................................................................165
6.1. The DHLLDV Framework. ................................................................................................................165
6.1.1. The Sliding Bed Regime. ...............................................................................................................165
6.1.2. The Heterogeneous Transport Regime. ..........................................................................................166
6.1.3. The Homogeneous Transport Regime. ...........................................................................................166
6.1.4. The Resulting Erhg Constant Spatial Volumetric Concentration Curve. .........................................168
6.1.5. Determining the Limit Deposit Velocity. .......................................................................................168
6.1.5.1. Introduction. ..........................................................................................................................168
6.1.5.2. Very Small & Small Particles. ...............................................................................................169
6.1.5.3. Large & Very Large Particles. ...............................................................................................169
6.1.5.4. The Resulting Upper Limit Froude Number..........................................................................170
6.1.5.5. The Lower Limit....................................................................................................................170
6.1.5.6. The Resulting Froude Number. .............................................................................................172
6.1.6. Nomenclature DHLLDV Framework. ............................................................................................173
6.2. The Jufin & Lopatin (1966) Model.....................................................................................................175
6.2.1. Introduction. ...................................................................................................................................175
6.2.2. Group A: Fines. ..............................................................................................................................175
6.2.3. Group B: Sand. ...............................................................................................................................175
6.2.4. The Limit Deposit Velocity. ...........................................................................................................178
6.2.5. Broad Graded Sands or Gravels. ....................................................................................................178
6.2.6. Group C: Fine Gravel. ....................................................................................................................179
6.2.7. Group D: Coarse Gravel. ................................................................................................................179
6.2.8. Conclusions & Discussion..............................................................................................................180
6.2.9. Nomenclature Early History & Empirical and Semi-Empirical Models. .......................................181
6.3. The Wilson-GIW (1992) Model for Heterogeneous Transport. .........................................................183
6.3.1. The Full Model. ..............................................................................................................................183
6.3.2. The Simplified Wilson Model. .......................................................................................................184
6.3.3. Generic Equation. ...........................................................................................................................184
6.3.4. Conclusions & Discussion..............................................................................................................185
6.3.5. Analysis of the v50 Equations. ........................................................................................................188
6.3.6. Near Wall Lift. ...............................................................................................................................192
6.3.7. The Demi-McDonald of Wilson (1979). ........................................................................................196
6.3.8. The Sliding Bed Regime New Developments. ...............................................................................197
6.3.9. Nomenclature Wilson-GIW Models...............................................................................................198
6.4. Notes. ..................................................................................................................................................201
Chapter 7: Modeling of the Swing Winches of a Cutter Dredge. ............................................................207
7.1. Introduction. .......................................................................................................................................207
7.2. The Motions of the Dredge. ................................................................................................................207

Copyright © Dr.ir. S.A. Miedema


Introduction Dredging Engineering.

7.3. The Influence of the Swing Angle on the Wire Moment. ...................................................................209
7.4. The Winch Characteristics. .................................................................................................................212
7.5. The Control System of the Winches. ..................................................................................................212
7.6. Case Studies........................................................................................................................................215
7.6.1. Case 1: ............................................................................................................................................215
7.6.2. Case 2: ............................................................................................................................................215
7.7. Conclusions. .......................................................................................................................................217
7.8. Nomenclature. ....................................................................................................................................218
7.9. Notes. ..................................................................................................................................................219
Chapter 8: The Trailing Suction Hopper Dredge. ....................................................................................225
8.1. Introduction ........................................................................................................................................225
8.2. The Loading Cycle of a Hopper Dredge .............................................................................................225
8.3. The Calculation Model. ......................................................................................................................230
8.4. The Layer Thickness of the Layer of Water above Overflow Level ..................................................231
8.5. The Storage Effect. .............................................................................................................................236
8.6. The Hopper of a TSHD as an Ideal Settlement Basin. ........................................................................237
8.7. The Modified Camp Model. ...............................................................................................................239
8.8. The Influence of Turbulence. .............................................................................................................245
8.9. Comparing the Miedema and the van Rhee Models. ..........................................................................249
8.9.1. Introduction. ...................................................................................................................................249
8.9.2. Case Studies with the Camp/Miedema Model. ..............................................................................249
8.9.3. The 2DV Model .............................................................................................................................252
8.9.4. Comparison of the Two Models. ....................................................................................................255
8.9.5. Conclusions ....................................................................................................................................258
8.10. A Sensitivity Analysis of the Scaling of TSHS’s. ..............................................................................259
8.10.1. Scale Laws. ................................................................................................................................259
8.10.2. The TSHD’S used. .....................................................................................................................260
8.10.3. Simulation Results. ....................................................................................................................261
8.10.4. Conclusions & Discussion. ........................................................................................................264
8.11. Waterjets in the Draghead. .................................................................................................................270
8.12. Conclusions & Discussion. .................................................................................................................273
8.13. Nomenclature .....................................................................................................................................275
8.14. Notes. ..................................................................................................................................................277
Chapter 9: References. ................................................................................................................................283
Chapter 10: List of Figures...........................................................................................................................299
Chapter 11: List of Tables. ...........................................................................................................................303

Copyright © Dr.ir. S.A. Miedema


Introduction Dredging Engineering.

Copyright © Dr.ir. S.A. Miedema


Basic Soil Mechanics.

Chapter 1: Basic Soil Mechanics.

1.1. Introduction.
Cutting processes of soil distinguish from the classical soil mechanics in civil engineering in the fact that:

Classical soil mechanics assume:


1. Small to very small strain rates.
2. Small to very small strains.
3. A very long time span, years to hundreds of years.
4. Structures are designed to last forever.
Cutting processes assume:
1. High to very high strain rates.
2. High to very high strains and deformations in general.
3. A very short time span, following from very high cutting velocities.
4. The soil is supposed to be excavated, the coherence has to be broken.

For the determination of cutting forces, power and specific energy the criterion for failure has to be known. In this
book the failure criterion of Mohr-Coulomb will be applied in the mathematical models for the cutting of sand,
clay and rock. The Mohr–Coulomb theory is named in honor of Charles-Augustin de Coulomb and Christian Otto
Mohr. Coulomb's contribution was a 1773 essay entitled "Essai sur une application des règles des maximis et
minimis à quelques problèmes de statique relatifs à l'architecture". Mohr developed a generalized form of the
theory around the end of the 19th century. To understand and work with the Mohr-Coulomb failure criterion it is
also necessary to understand the so called Mohr circle. The Mohr circle is a two dimensional graphical
representation of the state of stress at a point. The abscissa, σ, and ordinate, τ, of each point on the circle are the
normal stress and shear stress components, respectively, acting on a particular cut plane under an angle α with the
horizontal. In other words, the circumference of the circle is the locus of points that represent the state of stress on
individual planes at all their orientations. In this book a plane strain situation is considered, meaning a two-
dimensional cutting process. The width of the blades considered w is always much bigger than the layer thickness
hi considered. In geomechanics (soil mechanics and rock mechanics) compressive stresses are considered positive
and tensile stresses are considered to be negative, while in other engineering mechanics the tensile stresses are
considered to be positive and the compressive stresses are considered to be negative. Here the geomechanics
approach will be applied. There are two special stresses to be mentioned, the so called principal stresses. Principal
stresses occur at the planes where the shear stress is zero. In the plane strain situation there are two principal
stresses, which are always under an angle of 90º with each other.

In order to understand the cutting processes in sand, clay and rock, it is required to have knowledge of basic soil
and rock mechanics. The next chapters 1.2-1.7 cover this knowledge and have been composed almost entirely from
information from the public domain, especially internet. Most information comes from Wikipedia and
Answers.com.

1.2. Soil Mechanics.


1.2.1. Definition.
McGraw-Hill Science & Technology Encyclopedia gives the following description of Soil Mechanics:
The study of the response of masses composed of soil, water, and air to imposed loads. Because both water and
air are able to move through the soil pores, the discipline also involves the prediction of these transport processes.
Soil mechanics provides the analytical tools required for foundation engineering, retaining wall design, highway
and railway sub base design, tunneling, earth dam design, mine excavation design, and so on. Because the
discipline relates to rock as well as soils, it is also known as geotechnical engineering. Soil consists of a multiphase
aggregation of solid particles, water, and air.

This fundamental composition gives rise to unique engineering properties, and the description of the mechanical
behavior of soils requires some of the most sophisticated principles of engineering mechanics. The terms
multiphase and aggregation both imply unique properties. As a multiphase material, soil exhibits mechanical
properties that show the combined attributes of solids, liquids, and gases. Individual soil particles behave as solids,
and show relatively little deformation when subjected to either normal or shearing stresses. Water behaves as a
liquid, exhibiting little deformation under normal stresses, but deforming greatly when subjected to shear. Being

Copyright © Dr.ir. S.A. Miedema TOC Page 1 of 304


Introduction Dredging Engineering.

a viscous liquid, however, water exhibits a shear strain rate that is proportional to the shearing stress. Air in the
soil behaves as a gas, showing appreciable deformation under both normal and shear stresses. When the three
phases are combined to form a soil mass, characteristics that are an outgrowth of the interaction of the phases are
manifest. Moreover, the particulate nature of the solid particles contributes other unique attributes.

Figure 1-1: Earthwork in Germany (source Wikimedia).

When dry soil is subjected to a compressive normal stress, the volume decreases nonlinearly; that is, the more the
soil is compressed, the less compressible the mass becomes. Thus, the more tightly packed the particulate mass
becomes, the more it resists compression. The process, however, is only partially reversible, and when the
compressive stress is removed the soil does not expand back to its initial state.
When this dry particulate mass is subjected to shear stress, an especially interesting behavior owing to the
particulate nature of the soil solids results. If the soil is initially dense (tightly packed), the mass will expand
because the particles must roll up and over each other in order for shear deformation to occur. Conversely, if the
mass is initially loose, it will compress when subjected to a shear stress. Clearly, there must also exist a specific
initial density (the critical density) at which the material will display zero volume change when subjected to shear
stress. The term dilatancy has been applied to the relationship between shear stress and volume change in
particulate materials. Soil is capable of resisting shear stress up to a certain maximum value. Beyond this value,
however, the material undergoes large, uncontrolled shear deformation.
The other limiting case is saturated soil, that is, a soil whose voids are entirely filled with water. When such a mass
is initially loose and is subjected to compressive normal stress, it tends to decrease in volume; however, in order
for this volume decrease to occur, water must be squeezed from the soil pores. Because water exhibits a viscous
resistance to flow in the microscopic pores of fine-grained soils, this process can require considerable time, during
which the pore water is under increased pressure. This excess pore pressure is at a minimum near the drainage face
of the soil mass and at a maximum near the center of the soil sample. It is this gradient (or change in pore water
pressure with change in position within the soil mass) that causes the outflow of water and the corresponding
decrease in volume of the soil mass. Conversely, if an initially dense soil mass is subjected to shear stress, it tends
to expand. The expansion, however, may be time-dependent because of the viscous resistance to water being drawn
into the soil pores. During this time the pore water will be under decreased pressure. Thus, in saturated soil masses,
changes in pore water pressure and time-dependent volume change can be induced by either changes in normal
stress or by changes in shear stress.

Page 2 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Basic Soil Mechanics.

1.2.2. Soil Creation.


The primary mechanism of soil creation is the weathering of rock. All rock types (igneous rock, metamorphic rock
and sedimentary rock) may be broken down into small particles to create soil. Weathering mechanisms are physical
weathering, chemical weathering, and biological weathering. Human activities such as excavation, blasting, and
waste disposal, may also create soil. Over geologic time, deeply buried soils may be altered by pressure and
temperature to become metamorphic or sedimentary rock, and if melted and solidified again, they would complete
the geologic cycle by becoming igneous rock.
Physical weathering includes temperature effects, freeze and thaw of water in cracks, rain, wind, impact and other
mechanisms. Chemical weathering includes dissolution of matter composing a rock and composition of soils.
Physical weathering includes temperature effects, freeze and thaw of water in cracks, rain, wind, impact and other
mechanisms. Chemical weathering includes dissolution of matter composing a rock and precipitation in the form
of another mineral. Clay minerals, for example can be formed by weathering of feldspar, which is the most
common mineral present in igneous rock. The most common mineral constituent of silt and sand is quartz, also
called silica, which has the chemical name silicon dioxide. The reason that feldspar is most common in rocks but
silicon is more prevalent in soils is that feldspar is much more soluble than silica. Silt, Sand, and Gravel are
basically little pieces of broken rocks. According to the Unified Soil Classification System, silt particle sizes are
in the range of 0.002 mm to 0.075 mm and sand particles have sizes in the range of 0.075 mm to 4.75 mm. Gravel
particles are broken pieces of rock in the size range 4.75 mm to 100 mm. Particles larger than gravel are called
cobbles and boulders.

Figure 1-2: Fox glacier, New Zealand (source Wikimedia).

Soil deposits are affected by the mechanism of transport and deposition to their location. Soils that are not
transported are called residual soils -- they exist at the same location as the rock from which they were generated.
Decomposed granite is a common example of a residual soil. The common mechanisms of transport are the actions
of gravity, ice, water, and wind. Wind-blown soils include dune sands and loess. Water carries particles of different
size depending on the speed of the water, thus soils transported by water are graded according to their size. Silt
and clay may settle out in a lake, and gravel and sand collect at the bottom of a river bed. Wind-blown soil deposits
(aeolian soils) also tend to be sorted according to their grain size. Erosion at the base of glaciers is powerful enough

Copyright © Dr.ir. S.A. Miedema TOC Page 3 of 304


Introduction Dredging Engineering.

to pick up large rocks and boulders as well as soil; soils dropped by melting ice can be a well graded mixture of
widely varying particle sizes. Gravity on its own may also carry particles down from the top of a mountain to make
a pile of soil and boulders at the base; soil deposits transported by gravity are called colluvium.
The mechanism of transport also has a major effect on the particle shape. For example, low velocity grinding in a
river bed will produce rounded particles. Freshly fractured colluvium particles often have a very angular shape.

1.2.3. Soil Classification.


Soil classification deals with the systematic categorization of soils based on distinguishing characteristics as well
as criteria that dictate choices in use.

Figure 1-3: Soil naming according to USDA.

Soil texture is a qualitative classification tool used in both the field and laboratory to determine classes for
agricultural soils based on their physical texture. The classes are distinguished in the field by the 'textural feel'
which can be further clarified by separating the relative proportions of sand, silt and clay using grading sieves:
The Particle Size Distribution (PSD). The class is then used to determine crop suitability and to approximate the
soils responses to environmental and management conditions such as drought or calcium (lime) requirements. A
qualitative rather than a quantitative tool it is a fast, simple and effective means to assess the soils physical
characteristics. Although the U.S.D.A. system uses 12 classes whilst the U.K.-ADAS uses just 11 the systems are
mutually compatible as shown in the combined soil textural triangle below.

Page 4 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Basic Soil Mechanics.

Hand analysis, whilst an arbitrary technique, is an extremely simple and effective means to rapidly assess and
classify a soils physical condition. Correctly executed the procedure allows for rapid and frequent assessment of
soil characteristics with little or no equipment. It is thus an extremely useful tool for identifying spatial variation
both within and between plots (fields) as well as identifying progressive changes and boundaries between soil
classes and orders. The method involves taking a small sample of soil, sufficient to roll into a ball of approximately
2.5 cm diameter, from just below the surface. Using a small drop of water or 'spit' the sample is then moisten to
the sticky point (the point at which it begins to adhere to the finger). The ball is then molded to determine its
workability and its class according to the steps in the chart opposite.
Soil separates are specific ranges of particle sizes. In the United States, the smallest particles are clay particles and
are classified by the USDA as having diameters of less than 0.002 mm. The next smallest particles are silt particles
and have diameters between 0.002 mm and 0.05 mm. The largest particles are sand particles and are larger than
0.05 mm in diameter. Furthermore, large sand particles can be described as coarse, intermediate as medium, and
the smaller as fine. Other countries have their own particle size classifications.

Table 1-1: Soil Classification.

Name of Soil Diameter Limits (mm)


Clay <0.002
Fine silt 0.002–0.006
Medium silt 0.006-0.020
Coarse silt 0.020-0.060
Very fine sand 0.060–0.100
Fine sand 0.100–0.200
Medium sand 0.200–0.600
Coarse sand 0.600–1.000
Very coarse sand 1.000–2.000
Fine gravel 2-6
Medium gravel 6-20
Coarse gravel 20-60
Cobbles 60-200
Boulders >200

Figure 1-4: Soil failure (www.4isfge.org).

Copyright © Dr.ir. S.A. Miedema TOC Page 5 of 304


Introduction Dredging Engineering.

Figure 1-5: The Wenjiagou landslide (blogs.agu.org).

Figure 1-6: Karl von Terzaghi, one of the founders of modern soil mechanics.

Page 6 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Basic Soil Mechanics.

1.3. Soils.
1.3.1. Sand.
Sand is any material composed of loose, stony grains between 1/16 mm and 2 mm in diameter. Larger particles
are categorized as gravel; smaller particles are categorized as silt or clay. Sands are usually created by the
breakdown of rocks, and are transported by wind and water, before depositing to form soils, beaches, dunes, and
underwater fans or deltas. Deposits of sand are often cemented together over time to form sandstones.
The most common sand-forming process is weathering, especially of granite. Granite consists of distinct crystals
of quartz, feldspar, and other minerals. When exposed to water, some of these minerals (e.g., feldspar) decay
chemically faster than others (especially quartz), allowing the granite to crumble into fragments. Sand formed by
weathering is termed epiclastic.

Figure 1-7: Sand from the Gobi desert, Mongolia (source Wikimedia).

Where fragmentation is rapid, granite crumbles before its feldspar has fully decayed and the resulting sand contains
more feldspar. If fragmentation is slow, the resulting sand contains less feldspar. Fragmentation of rock is enhanced
by exposure to fast-running water, so steep mountains are often source areas for feldspar-rich sands and gentler
terrains are often source areas for feldspar-poor sands. Epiclastic sands and the sandstones formed from them thus
record information about the environments that produce them. A sedimentologist can deduce the existence of
whole mountain ranges long ago eroded, and of mountain-building episodes that occurred millions of years ago
from sandstones rich in relatively unstable minerals like feldspar.
The behavior of sand carried by flowing water can inscribe even more detailed information about the environment
in sand deposits. When water is flowing rapidly over a horizontal surface, any sudden vertical drop in that surface
splits the current into two layers, (1) an upper layer that continues to flow downstream and (2) a slower backflow
that curls under in the lee of the drop-off. Suspended sand tends to settle out in the backflow zone, building a slope
called a "slip face" that tilts downhill from the drop-off. The backflow zone adds continually to the slip face,
growing it downstream, and as the slip face grows downstream its top edge continues to create a backflow zone.
The result is the deposition of a lengthening bed of sand. Typically, periodic avalanches of large grains down the
slip face (or other processes) coat it with thin layers of distinctive material. These closely-spaced laminations are
called "cross bedding" because they angle across the main bed. Cross-bedding in sandstone records the direction
of the current that deposited the bed, enabling geologists to map currents that flowed millions of years ago
(paleocurrents).
Evidence of grain size, bed thickness, and cross-bedding angle, allows geologists to determine how deep and fast
a paleocurrent was, and thus how steep the land was over which it flowed.

Copyright © Dr.ir. S.A. Miedema TOC Page 7 of 304


Introduction Dredging Engineering.

Figure 1-8: Sand in the Sahara desert (source Luca Galuzzi – www.galuzzi.it)

Ripples and dunes—probably the most familiar forms created by wind- or waterborne sand—involve similar
processes. However, ripples and dunes are more typical of flow systems to which little or no sand is being added.
The downstream slip faces of ripples and dunes are built from grains plucked from their upstream sides, so these
structures can migrate without growing. When water or wind entering the system (e.g., water descending rapidly
from a mountainous region) imports large quantities of sand, the result is net deposition rather than the mere
migration of sand forms.
Grain shape, too, records history. All epiclastic grains of sand start out angular and become more rounded as they
are polished by abrasion during transport by wind or water. Quartz grains, however, resist wear. One trip down a
river is not enough to thoroughly round an angular grain of quartz; even a long sojourn on a beach, where grains
are repeatedly tumbled by waves, does not suffice. The well-rounded state of many quartz sands can be accounted
for only by crustal recycling. Quartz grains can survive many cycles of erosion, burial and cementation into
sandstone, uplift, and re-erosion. Recycling time is on the order of 200 million years, so a quartz grain first
weathered from granite 2.4 billion years ago may have gone through 10 or 12 cycles of burial and re-erosion to
reach its present day state. An individual quartz grain's degree of roundness is thus an index of its antiquity.
Feldspar grains can also survive recycling, but not as well, so sand that has been recycled a few times consists
mostly of quartz.
Sand can be formed not only by weathering but by explosive volcanism, the breaking up of shells by waves, the
cementing into pellets of finer-grained materials (pelletization), and the precipitation of dissolved chemicals (e.g.,
calcium carbonate) from solution.
Pure quartz sands are mined to make glass and the extremely pure silicon employed in microchips and other
electronic components.

Page 8 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Basic Soil Mechanics.

1.3.2. Clay.
Clay is a fine-grained (small particle size) sedimentary rock. Clay is so fine-grained it is rarely possible to see the
individual mineral particles with the naked eye. The definition of clays describes rocks with particle sizes of less
than 4 μm in diameter. Most sedimentary rocks are described using both mineral content and particle size. While
this is also true for clays, the particle size description is most reliable and most often used.

Figure 1-9: Quaternary clay in Estonia (source Wikimedia)

The majority of common types of minerals found in clays are kaolinite (a soapy-feeling and lightweight mineral),
talc, pyrophyllite, all types of micas, minerals from the chlorite group, feldspars, and a lesser amount of
tectosilicates (including quartz).
The mineral content of clays is less variable than other types of sedimentary rock. This is a direct result of the way
clays are formed. Water carries the bulk of sediments to their resting place where they are cemented together. The
transport of sediments is directly related to the force or velocity of water carrying them. The stronger the velocity
of water, the larger and heavier the particle it can move. Conversely, the weaker the flow, the smaller the particle
that is carried by the water. As a result, water acts as a winnowing filter for certain types of minerals. The heavier
minerals are not carried as far by water currents as are the lighter ones. When water finally comes to rest, it deposits
its load of minerals. The last to be released are the lighter and smaller particles, the clay minerals.
Where rivers meet oceans, the clay minerals are so light they are usually carried far out to sea where they fall
gently to the bottom forming a fine-grained sediment. These deposits cover organic materials and trap them at the
edges of deltas and continental slopes. Over millions of years, the organic materials convert to petroleum and
remain trapped by the clays. This relationship makes the study of clays extremely important for petroleum
geologists. In addition to this important economic consideration, clays provide important economic resources for
a wide variety of other industries.

Depending on the academic source, there are three or four main groups of clays: kaolinite, montmorillonite,
smectite, illite, and chlorite. Chlorites are not always considered a clay, sometimes being classified as a separate
group within the phyllosilicates. There are approximately 30 different types of "pure" clays in these categories,
but most "natural" clays are mixtures of these different types, along with other weathered minerals.

Copyright © Dr.ir. S.A. Miedema TOC Page 9 of 304


Introduction Dredging Engineering.

Varve (or varved clay) is clay with visible annual layers, formed by seasonal differences in erosion and organic
content. This type of deposit is common in former glacial lakes. When glacial lakes are formed there is very little
movement of the water that makes the lake, and these eroded soils settle on the lake bed. This allows such an even
distribution on the different layers of clay.

Figure 1-10: Varved clay, Little River State Park, Waterbury, Vermont
(source www.anr.state.vt.us).

Quick clay is a unique type of marine clay indigenous to the glaciated terrains of Norway, Canada, Northern
Ireland, and Sweden. It is highly sensitive clay, prone to liquefaction, which has been involved in several deadly
landslides.
Clays exhibit plasticity when mixed with water in certain proportions. When dry, clay becomes firm and when
fired in a kiln, permanent physical and chemical changes occur. These reactions, among other changes, cause the
clay to be converted into a ceramic material. Because of these properties, clay is used for making pottery items,
both utilitarian and decorative. Different types of clay, when used with different minerals and firing conditions,
are used to produce earthenware, stoneware, and porcelain. Prehistoric humans discovered the useful properties of
clay, and one of the earliest artifacts ever uncovered is a drinking vessel made of sun-dried clay. Depending on the
content of the soil, clay can appear in various colors, from a dull gray to a deep orange-red.
Clay tablets were used as the first known writing medium, inscribed with cuneiform script through the use of a
blunt reed called a stylus.
Clays sintered in fire were the first form of ceramic. Bricks, cooking pots, art objects, dishware, and even musical
instruments such as the ocarina can all be shaped from clay before being fired. Clay is also used in many industrial
processes, such as paper making, cement production, and chemical filtering. Clay is also often used in the
manufacture of pipes for smoking tobacco. Until the late 20th century bentonite clay was widely used as a mold
binder in the manufacture of sand castings.
Clay, being relatively impermeable to water, is also used where natural seals are needed, such as in the cores of
dams, or as a barrier in landfills against toxic seepage (lining the landfill, preferably in combination with
geotextiles).
Recent studies have investigated clay's absorption capacities in various applications, such as the removal of heavy
metals from waste water and air purification.

Page 10 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Basic Soil Mechanics.

1.3.3. Rock.
To the geologist, the term rock means a naturally occurring aggregate of minerals that may include some organic
solids (e.g., fossils) and/or glass. Rocks are generally subdivided into three large classes: igneous, sedimentary,
and metamorphic. These classes relate to common origin, or genesis. Igneous rocks form from cooling liquid rock
or related volcanic eruptive processes. Sedimentary rocks form from compaction and cementation of sediments.
Metamorphic rocks develop due to solid-state, chemical and physical changes in pre-existing rock because of
elevated temperature, pressure, or chemically active fluids.
With igneous rocks, the aggregate of minerals comprising these rocks forms upon cooling and crystallization of
liquid rock. As crystals form in the liquid rock, they become interconnected to one another like jigsaw puzzle
pieces. After total crystallization of the liquid, a hard, dense igneous rock is the result. Also, some volcanic lavas,
when extruded on the surface and cooled instantaneously, will form a natural glass.

Figure 1-11: Sample of igneous gabbro, Rock Creek Canyon, California (source Wikimedia).

Glass is a mass of disordered atoms, which are frozen in place due to sudden cooling, and is not a crystalline
material like a mineral. Glass composes part of many extrusive igneous rocks (e.g., lava flows) and pyroclastic
igneous rocks. Alternatively, some igneous rocks are formed from volcanic processes, such as violent volcanic
eruption. Violent eruptions eject molten, partially molten, and non-molten igneous rock, which then falls in the
vicinity of the eruption. The fallen material may solidify into a hard mass, called pyroclastic igneous rock. The
texture of igneous rocks (defined as the size of crystals in the rock) is strongly related to cooling rate of the original
liquid. Rapid cooling of liquid rock promotes formation of small crystals, usually too small to see with the unaided
eye. Rocks with this cooling history are called fine-textured igneous rocks. Slow cooling (which usually occurs
deep underground) promotes formation of large crystals. Rocks with this cooling history are referred to as coarse-
textured igneous rocks.
The mineral composition of igneous rocks falls roughly into four groups: silicic, intermediate, mafic, and
ultramafic. These groups are distinguished by the amount of silica (SiO 4), iron (Fe), and magnesium (Mg) in the
constituent minerals. Mineral composition of liquid rock is related to place of origin within the body of the earth.
Generally speaking, liquids from greater depths within the earth contain more Fe and Mg and less SiO4 than those
from shallow depths.
In sedimentary rocks, the type of sediment that is compacted and cemented together determines the rock's main
characteristics. Sedimentary rocks composed of sediment that has been broken into pieces (i.e., clastic sediment),

Copyright © Dr.ir. S.A. Miedema TOC Page 11 of 304


Introduction Dredging Engineering.

such as gravel, sand, silt, and clay, are clastic sedimentary rocks (e.g., conglomerate, sandstone, siltstone, and
shale, respectively). Sedimentary rocks composed of sediment that is chemically derived (i.e., chemical sediment),
such as dissolved elements like calcium (Ca), sodium (Na), iron (Fe), and silicon (Si), are chemical sedimentary
rocks. Examples of chemical sedimentary rocks are limestone (composed of calcium carbonate), rock salt
(composed of sodium chloride), rock gypsum (composed of calcium sulfate), ironstones (composed of iron oxides),
and chert (composed of hydrated silica). Biochemical sedimentary rocks are a special kind of chemical sedimentary
rock wherein the constituent particles were formed by organisms (typically as organic hard parts, such as shells),
which then became sedimentary particles. Examples of this special kind of sedimentary rock include chalk,
fossiliferous limestone, and coquina. Sedimentary rocks are formed from sediment in two stages: compaction and
cementation. Compaction occurs when sediments pile up to sufficient thickness that overlying mass squeezes out
water and closes much open space. Cementation occurs when water flowing through the compacted sediment
deposits mineral crystals upon particles thus binding them together. The main cement minerals are calcite (CaCO 3),
hematite (Fe2O3), and quartz (SiO2).
With metamorphic rocks, the nature of the pre-existing rock (protolith) determines in large part the characteristics
of the ultimate metamorphic rock. Regardless of protolith, however, almost all metamorphic rocks are harder and
more dense than their protoliths. A protolith with flat or elongate mineral crystals (e.g., micas or amphiboles) will
yield a metamorphic rock with preferentially aligned minerals (due to directed pressure). Such metamorphic rocks
are called foliated metamorphic rocks (e.g., slate and schist). Non-foliated metamorphic rocks (e.g., marble and
quartzite) come from protoliths that have mainly equidimensional mineral crystals (e.g., calcite and quartz,
respectively). For example, a protolith shale will yield a foliated metamorphic rock, and a protolith limestone will
yield marble, a non-foliated metamorphic rock. Metamorphic rocks possess distinctive grades or levels of
metamorphic change from minimal to a maximum near total melting. Low-grade metamorphic rocks generally
have fine-textured crystals and low-temperature indicator minerals like the mica chlorite. High-grade metamorphic
rocks generally have coarse-textured crystals and very distinctive foliation, plus high-temperature indicator
minerals like the silicate mineral staurolite.
Rock is a brittle natural solid found mainly in the outer reaches of Earth's crust and upper mantle. Material that
would be brittle rock at such shallow depths becomes to one degree or another rather plastic within the body of the
earth. The term "rock" is not generally applied to such non-brittle internal Earth materials. Therefore, rock is a
concept related to the outer shell of the earth. The term rock may also be properly applied to brittle natural solids
found on the surfaces of other planets and satellites in our solar system. Meteorites are rock. Naturally occurring
ice (e.g., brittle water ice in a glacier, H2O) is also a rock, although we do not normally think of ice this way.
Rock has been an important natural resource for people from early in human evolution. Rocks' properties are the
key to their specific usefulness, now as in the past. Hard, dense rocks that could be chipped into implements and
weapons were among the first useful possessions of people. Fine-textured and glassy rocks were particularly handy
for these applications. Later on, rock as building stone and pavement material became very important, and this
continues today in our modern world. All of Earth's natural mineral wealth, fossil energy resources, and most
groundwater are contained within rocks of the earth's crust.

Rock is a natural occurrence mass of cohesive organic or inorganic material, which forms a part earth crest of
which most rocks are composed of one or more minerals. Rocks can be classified in different ways. The most used
classification is based on their origin, in which the following classes can be distinguished.
Igneous rock; a rock that has solidified from molten rock material (magma), which was generated within the Earth.
Well known are granite and basalt
Sedimentary rock; a rock formed by the consolidation of sediment settle out in water, ice of air and accumulated
on the Earth’s surface, either on dry land or under water. Examples are sandstone, lime stone and clay stone
Metamorphic rock; any class of rocks that are the result of partial or complete recrystallization in the solid state of
pre-existing rocks under conditions of temperature and pressure that are significantly different from those
obtaining at the surface of the Earth.
When deterring the dredge-ability of rock, distinction has to be made between the properties of intact rock and that
of a rock mass. Depending on the fracture density of the rock the cutter will cut intact rock or break out rock
blocks.
In the first case the strength (tensile- and compressive strength), deformation properties (E-value) and the
petrography (mineralogical proposition) of the intact rock determines the production completely. The second case
the fracture frequency and the weathering of the rock is more important than the strength of the intact rock. It is
known that the absence of water in rock is important for the rock strength. When saturated with water the rock
strength can be 30 to 90 % of the strength of dry rock. Therefore rock samples have to be sealed immediately after
drilling in such a way that evaporation of or intake of water is avoided. It has to be mentioned that this does not
mean that cutting forces in saturated rock are always lower than in dry rock. The petrography is important for the
weir of rock cutting tools.

Page 12 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Basic Soil Mechanics.

Figure 1-12: Sandstone formations, Vermillion Cliffs, Arizona (source www.reddit.com).

Figure 1-13: Columns of Basalt of the Scottish Island of Staffa (National Geographic).

Copyright © Dr.ir. S.A. Miedema TOC Page 13 of 304


Introduction Dredging Engineering.

Figure 1-14 A: Aid to identification of rock for engineering purposes


(After BS 5930:1981).

Page 14 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Basic Soil Mechanics.

Figure 1-15 B: Aid to identification of rock for engineering purposes


(After BS 5930:1981).

Copyright © Dr.ir. S.A. Miedema TOC Page 15 of 304


Introduction Dredging Engineering.

Figure 1-16: Utica Shale, Fort Plain, New York (Wikipedia).

Figure 1-17: The rock formation cycle (galleryhip.com).

Page 16 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Basic Soil Mechanics.

1.4. Soil Mechanical Parameters.


1.4.1. Grain Size Distribution/Particle Size Distribution.
Soils consist of a mixture of particles of different size, shape and mineralogy. Because the size of the particles
obviously has a significant effect on the soil behavior, the grain size and grain size distribution are used to classify
soils. The grain size distribution describes the relative proportions of particles of various sizes. The grain size is
often visualized in a cumulative distribution graph which, for example, plots the percentage of particles finer than
a given size as a function of size. The median grain size, d50, is the size for which 50% of the particle mass consists
of finer particles. Soil behavior, especially the hydraulic conductivity, tends to be dominated by the smaller
particles; hence, the term "effective size", denoted by d10, is defined as the size for which 10% of the particle mass
consists of finer particles.
Sands and gravels that possess a wide range of particle sizes with a smooth distribution of particle sizes are called
well graded soils. If the soil particles in a sample are predominantly in a relatively narrow range of sizes, the soil
is called uniformly graded soils. If there are distinct gaps in the gradation curve, e.g., a mixture of gravel and fine
sand, with no coarse sand, the soils may be called gap graded. Uniformly graded and gap graded soils are both
considered to be poorly graded. There are many methods for measuring particle size distribution. The two
traditional methods used in geotechnical engineering are sieve analysis and hydrometer analysis .

Cumulative Grain Size Distribution, Data Roberts et al. (1998)


100
90
80
% Finer by Weight

70
60
50
40
30
20
10
0
10-4 10-3 10-2 10-1 100 101
Grain Size in mm
d50=0.0057 mm d50=0.0148 mm d50=0.0183 mm d50=0.0480 mm d50=0.0750 mm

d50=0.0125 mm d50=0.2220 mm d50=0.4320 mm d50=1.0200 mm d50=1.3500 mm

d50=0.0057 mm d50=0.0148 mm d50=0.0183 mm d50=0.0480 mm d50=0.0750 mm

d50=0.125 mm d50=0.2220 mm d50=0.4320 mm d50=1.0200 mm d50=1.3500 mm

Figure 1-18: The particle size distributions of the sands used by


Roberts et al. (1998).

1.4.2. Atterberg Limits.


The Atterberg limits are a basic measure of the nature of a fine-grained soil. Depending on the water content of
the soil, it may appear in four states: solid, semi-solid, plastic and liquid. In each state the consistency and behavior
of a soil is different and thus so are its engineering properties. Thus, the boundary between each state can be
defined based on a change in the soil's behavior. The Atterberg limits can be used to distinguish between silt and
clay, and it can distinguish between different types of silts and clays. These limits were created by Albert Atterberg,
a Swedish chemist. They were later refined by Arthur Casagrande. These distinctions in soil are used in picking
the soils to build structures on top of. These tests are mainly used on clayey or silty soils since these are the soils
that expand and shrink due to moisture content. Clays and silts react with the water and thus change sizes and have
varying shear strengths. Thus these tests are used widely in the preliminary stages of building any structure to

Copyright © Dr.ir. S.A. Miedema TOC Page 17 of 304


Introduction Dredging Engineering.

insure that the soil will have the correct amount of shear strength and not too much change in volume as it expands
and shrinks with different moisture contents.

Figure 1-19: Liquid limit device. Figure 1-20: Liquid limit device.

1.4.2.1. Shrinkage Limit.


The shrinkage limit (SL) is the water content where further loss of moisture will not result in any more volume
reduction. The test to determine the shrinkage limit is ASTM International D4943. The shrinkage limit is much
less commonly used than the liquid and plastic limits.

1.4.2.2. Plastic Limit.


The plastic limit (PL) is the water content where soil transitions between brittle and plastic behavior. A thread of
soil is at its plastic limit when it begins to crumble when rolled to a diameter of 3 mm. To improve test result
consistency, a 3 mm diameter rod is often used to gauge the thickness of the thread when conducting the test. The
Plastic Limit test is defined by ASTM standard test method D 4318.

1.4.2.3. Liquid Limit.


The liquid limit (LL) is the water content at which a soil changes from plastic to liquid behavior. The original
liquid limit test of Atterberg's involved mixing a pat of clay in a round-bottomed porcelain bowl of 10-12cm
diameter. A groove was cut through the pat of clay with a spatula, and the bowl was then struck many times against
the palm of one hand. Casagrande subsequently standardized the apparatus and the procedures to make the
measurement more repeatable. Soil is placed into the metal cup portion of the device and a groove is made down
its center with a standardized tool of 13.5 millimeters (0.53 in) width. The cup is repeatedly dropped 10mm onto
a hard rubber base during which the groove closes up gradually as a result of the impact. The number of blows for
the groove to close is recorded. The moisture content at which it takes 25 drops of the cup to cause the groove to
close over a distance of 13.5 millimeters (0.53 in) is defined as the liquid limit. The test is normally run at several
moisture contents, and the moisture content which requires 25 blows to close the groove is interpolated from the
test results. The Liquid Limit test is defined by ASTM standard test method D 4318. The test method also allows
running the test at one moisture content where 20 to 30 blows are required to close the groove; then a correction
factor is applied to obtain the liquid limit from the moisture content.
The following is when you should record the N in number of blows needed to close this 1/2-inch gap:
The materials needed to do a Liquid limit test are as follows
 Casagrande cup ( liquid limit device)
 Grooving tool
 Soil pat before test
 Soil pat after test
Another method for measuring the liquid limit is the fall cone test. It is based on the measurement of penetration
into the soil of a standardized cone of specific mass. Despite the universal prevalence of the Casagrande method,
the fall cone test is often considered to be a more consistent alternative because it minimizes the possibility of
human variations when carrying out the test.

Page 18 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Basic Soil Mechanics.

1.4.2.4. Importance of Liquid Limit Test.


The importance of the liquid limit test is to classify soils. Different soils have varying liquid limits. Also to find
the plasticity index of a soil you need to know the liquid limit and the plastic limit.

1.4.2.5. Derived Limits.


The values of these limits are used in a number of ways. There is also a close relationship between the limits and
properties of a soil such as compressibility, permeability, and strength. This is thought to be very useful because
as limit determination is relatively simple, it is more difficult to determine these other properties. Thus the
Atterberg limits are not only used to identify the soil's classification, but it allows for the use of empirical
correlations for some other engineering properties.

1.4.2.6. Plasticity Index.


The plasticity index (PI) is a measure of the plasticity of a soil. The plasticity index is the size of the range of water
contents where the soil exhibits plastic properties. The PI is the difference between the liquid limit and the plastic
limit (PI = LL-PL). Soils with a high PI tend to be clay, those with a lower PI tend to be silt, and those with a PI
of 0 (non-plastic) tend to have little or no silt or clay.
PI and their meanings
 0 – Non-plastic
 (1-5)- Slightly Plastic
 (5-10) - Low plasticity
 (10-20)- Medium plasticity
 (20-40)- High plasticity
 >40 Very high plasticity

1.4.2.7. Liquidity Index.


The liquidity index (LI) is used for scaling the natural water content of a soil sample to the limits. It can be
calculated as a ratio of difference between natural water content, plastic limit, and plasticity index:
LI=(W-PL)/(LL-PL) where W is the natural water content.

1.4.2.8. Activity.
The activity (A) of a soil is the PI divided by the percent of clay-sized particles (less than 2 μm) present. Different
types of clays have different specific surface areas which controls how much wetting is required to move a soil
from one phase to another such as across the liquid limit or the plastic limit. From the activity one can predict the
dominant clay type present in a soil sample. High activity signifies large volume change when wetted and large
shrinkage when dried. Soils with high activity are very reactive chemically. Normally the activity of clay is
between 0.75 and 1.25, and in this range clay is called normal. It is assumed that the plasticity index is
approximately equal to the clay fraction (A = 1). When A is less than 0.75, it is considered inactive. When it is
greater than 1.25, it is considered active.

Figure 1-21: The relation between SL, PL, LL and PI.

Copyright © Dr.ir. S.A. Miedema TOC Page 19 of 304


Introduction Dredging Engineering.

1.4.3. Mass Volume Relations.


There are a variety of parameters used to describe the relative proportions of air (gas), water (liquid) and solids in
a soil. This section defines these parameters and some of their interrelationships. The basic notation is as follows:
Vg, Vl, and Vs represent the volumes of gas, liquid and solids in a soil mixture;
Wg, Wl, and Ws represent the weights of gas, liquid and solids in a soil mixture;
Mg, Ml, and Ms represent the masses of gas, liquid and solids in a soil mixture;
ρg, ρl, and ρs represent the densities of the constituents (gas, liquid and solids) in a soil mixture;
Note that the weights, W, can be obtained by multiplying the mass, M, by the acceleration due to gravity, g; e.g.,
Ws = Ms·g

1.4.3.1. Specific Gravity.


Specific Gravity is the ratio of the density of one material compared to the density of pure water (ρl = 1000 kg/m3).

s
G s  (1-1)
l

1.4.3.2. Density.
The terms density and unit weight are used interchangeably in soil mechanics. Though not critical, it is important
that we know it. Density, Bulk Density, or Wet Density, ρt, are different names for the density of the mixture, i.e.,
the total mass of air, water, solids divided by the total volume of air, water and solids (the mass of air is assumed
to be zero for practical purposes. To find the formula for density, divide the mass of the soil by the volume of the
soil, the basic formula for density is:

M M s  M l  M g
t
t   (1-2)
Vt Vs  Vl  Vg

Unit weight of a soil mass is the ratio of the total weight of soil to the total volume of soil. Unit Weight, t, is
usually determined in the laboratory by measuring the weight and volume of a relatively undisturbed soil sample
obtained from a brass ring. Measuring unit weight of soil in the field may consist of a sand cone test, rubber balloon
or nuclear densitometer, the basic formula for unit weight is:

M t g
t  (1-3)
Vt

Dry Density, ρd, is the mass of solids divided by the total volume of air, water and solids:

M s M s
d   (1-4)
Vt Vs  Vl  Vg

Submerged Density, ρst, defined as the density of the mixture minus the density of water is useful if the soil is
submerged under water:

 sd   t   l (1-5)

Page 20 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Basic Soil Mechanics.

Table 1-2: Empirical values for ρt, of granular soils based on the standard penetration number, (from
Bowels, Foundation Analysis).

SPT Penetration, N-Value ρt (kg/m3)


(blows/ foot)
0-4 1120 - 1520
4 - 10 1520 - 1800
10 - 30 1800 - 2080
30 - 50 2080 - 2240
>50 2240 - 2400

Table 1-3: Empirical values for ρs, of cohesive soils based on the standard penetration number, (From
Bowels, Foundation Analysis).

SPT Penetration, N-Value (blows/ foot) ρs, sat (kg/m3)


0-4 1600 - 1840
4-8 1840 - 2000
8 - 32 2000 - 2240

Table 1-4: Typical Soil Characteristics (From Lindeburg, Civil Engineering Reference Manual for the PE
Exam, 8th edition).

Soil Type ρs (kg/m3) ρs, sat (kg/m3)


Sand, loose and uniform 1440 1888
Sand, dense and uniform 1744 2080
Sand, loose and well graded 1584 1984
Sand, dense and well graded 1856 2160
Glacial clay, soft 1216 1760
Glacial clay, stiff 1696 2000

Table 1-5: Typical Values of Soil Index Properties


(From NAVFAC 7.01).

Soil Type ρs (kg/m3)


Sand; clean, uniform, fine or medium 1344 - 2176
Silt; uniform, inorganic 1296 - 2176
Silty Sand 1408 - 2272
Sand; Well-graded 1376 - 2368
Silty Sand and Gravel 1440 - 2480
Sandy or Silty Clay 1600 - 2352
Silty Clay with Gravel; uniform 1840 - 2416
Well-graded Gravel, Sand, Silt and Clay 2000 - 2496
Clay 1504 - 2128
Colloidal Clay 1136 - 2048
Organic Silt 1392 - 2096
Organic Clay 1296 - 2000

1.4.3.3. Relative Density.


Relative density is an index that quantifies the state of compactness between the loosest and densest possible state
of coarse-grained soils. The relative density is written in the following formulas:

Copyright © Dr.ir. S.A. Miedema TOC Page 21 of 304


Introduction Dredging Engineering.

em ax  e n m ax  n
Dr   (1-6)
e m a x  e m in n m a x  n m in

Table 1-6: Designation of Granular Soil Based on Relative Density.

Dr (%) Description
0 - 20 Very loose
20 - 40 Loose
40 - 70 Medium dense
70 - 85 Dense
85 - 100 Very dense

SPT values versus relative density.


100.0

90.0

80.0
SPT value in blows/305 mm

70.0

60.0

50.0

40.0

30.0

20.0

10.0

0.0
0 10 20 30 40 50 60 70 80 90 100
Relative density in %
0 kPa 69 kPa 138 kPa 276 kPa

z = 0m z = 10 m z = 20 m z = 30 m

Figure 1-22: SPT values versus relative density (Miedema (1995).

Lambe & Whitman (1979), page 78 (Figure 1-22) give the relation between the SPT value, the relative density and
the hydrostatic pressure in two graphs. With some curve-fitting these graphs can be summarized with the following
equation (Miedema (1995)):

S P T   1 .8 2  0 .2 2 1   z  1 0    1 0
4 2 .5 2
 RD (1-7)

Page 22 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Basic Soil Mechanics.

1.4.3.4. Porosity.
Porosity is the ratio of the volume of openings (voids) to the total volume of material. Porosity represents the
storage capacity of the geologic material. The primary porosity of a sediment or rock consists of the spaces between
the grains that make up that material. The more tightly packed the grains are, the lower the porosity. Using a box
of marbles as an example, the internal dimensions of the box would represent the volume of the sample. The space
surrounding each of the spherical marbles represents the void space. The porosity of the box of marbles would be
determined by dividing the total void space by the total volume of the sample and expressed as a percentage.
The primary porosity of unconsolidated sediments is determined by the shape of the grains and the range of grain
sizes present. In poorly sorted sediments, those with a larger range of grain sizes, the finer grains tend to fill the
spaces between the larger grains, resulting in lower porosity. Primary porosity can range from less than one percent
in crystalline rocks like granite to over 55% in some soils. The porosity of some rock is increased through fractures
or solution of the material itself. This is known as secondary porosity.

Vv Vv e
n    (1-8)
Vt Vs  V v 1 e

1.4.3.5. Void ratio.


The ratio of the volume of void space to the volume of solid substance in any material consisting of void space
and solid material, such as a soil sample, a sediment, or a powder.

Vv Vv n
e    (1-9)
Vs Vt  Vv 1 n

The relations between void ratio e and porosity n are:

n e
e  and n= (1-10)
1n 1 e

1.4.3.6. Dilatation.
Dilation (or dilatation) refers to an enlargement or expansion in bulk or extent, the opposite of contraction. It
derives from the Latin dilatare, "to spread wide". It is the increase in volume of a granular substance when its shape
is changed, because of greater distance between its component particles. Suppose we have a volume V before the
enlargement and a volume V+dV after the enlargement. Before the enlargement we name the porosity ni (i from
initial) and after the enlargement ncv (the constant volume situation after large deformations). For the volume
before the deformation we can write:

V  1  ni   V  ni  V (1-11)

The first term on the right hand side is the sand volume, the second term the pore volume. After the enlargement
we get:

V  dV   1  n cv    V  dV  n cv   V  d V  (1-12)

Again the first term on the right hand side is the sand volume. Since the sand volume did not change during the
enlargement (we assume the quarts grains are incompressible), the volume of sand in both equations should be the
same, thus:

1  ni   V   1  n cv    V  dV  (1-13)

From this we can deduce that the dilatation ε is:

dV n cv  n i dn
    (1-14)
V 1  n cv 1  n cv

Copyright © Dr.ir. S.A. Miedema TOC Page 23 of 304


Introduction Dredging Engineering.

1.4.4. Permeability.
Permeability is a measure of the ease with which fluids will flow though a porous rock, sediment, or soil. Just as
with porosity, the packing, shape, and sorting of granular materials control their permeability. Although a rock
may be highly porous, if the voids are not interconnected, then fluids within the closed, isolated pores cannot
move. The degree to which pores within the material are interconnected is known as effective porosity. Rocks such
as pumice and shale can have high porosity, yet can be nearly impermeable due to the poorly interconnected voids.
In contrast, well-sorted sandstone closely replicates the example of a box of marbles cited above. The rounded
sand grains provide ample, unrestricted void spaces that are free from smaller grains and are very well linked.
Consequently, sandstones of this type have both high porosity and high permeability.
The range of values for permeability in geologic materials is extremely large. The most conductive materials have
permeability values that are millions of times greater than the least permeable. Permeability is often directional in
nature. The characteristics of the interstices of certain materials may cause the permeability to be significantly
greater in one direction. Secondary porosity features, like fractures, frequently have significant impact on the
permeability of the material. In addition to the characteristics of the host material, the viscosity and pressure of the
fluid also affect the rate at which the fluid will flow.

Hydraulic conductivity or permeability k can be estimated by particle size analysis of the sediment of interest,
using empirical equations relating either k to some size property of the sediment. Vukovic and Soro (1992)
summarized several empirical methods from former studies and presented a general formula:

g
f n  de
2
k  C (1-15)
l
The kinematic viscosity vl is related to dynamic viscosity µl and the fluid (water) density ρl as follows:

l
l  (1-16)
l

The values of C, f(n) and de are dependent on the different methods used in the grain-size analysis. According to
Vukovic and Soro (1992), porosity n may be derived from the empirical relationship with the coefficient of grain
uniformity U as follows:


n  0 .2 5 5  1  0 .8 3
U
 (1-17)

Where U is the coefficient of grain uniformity and is given by:

 d 60 
U    (1-18)
 d 10 

Here, d60 and d10 in the formula represent the grain diameter in (mm) for which, 60% and 10% of the sample
respectively, are finer than. Former studies have presented the following formulae which take the general form
presented in equation (1-15) above but with varying C, f(n) and de values and their domains of applicability.

Hazen’s formula (1982) was originally developed for determination of hydraulic conductivity of uniformly graded
sand but is also useful for fine sand to gravel range, provided the sediment has a uniformity coefficient less than 5
and effective grain size between 0.1 and 3mm.

g
  1  1 0   n  0 .2 6    d 1 0
4 2
k  6  10  (1-19)
l

The Kozeny-Carman equation is one of the most widely accepted and used derivations of permeability as a function
of the characteristics of the soil medium. The Kozeny-Carman equation (or Carman-Kozeny equation) is a relation
used in the field of fluid dynamics to calculate the pressure drop of a fluid flowing through a packed bed of solids.
It is named after Josef Kozeny and Philip C. Carman. This equation was originally proposed by Kozeny (1927)
and was then modified by Carman (1937) and (1956) to become the Kozeny-Carman equation. It is not appropriate

Page 24 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Basic Soil Mechanics.

for either soil with effective size above 3 mm or for clayey soils. The equation is only valid for laminar flow. The
equation is given as:

3 3
2 l e 3 g n 2
k  de   C or k = 8 .3  1 0    d 10
l 1 e l 1  n 
2

(1-20)
l
W it h : l  and  l   l  g
l

This equation holds for flow through packed beds with particle Reynolds numbers up to approximately 1.0, after
which point frequent shifting of flow channels in the bed causes considerable kinetic energy losses. This equation
can be expressed as "flow is proportional to the pressure drop and inversely proportional to the fluid viscosity",
which is known as Darcy's law.

The Breyer method does not consider porosity and therefore, porosity function takes on value 1. Breyer formula
is often considered most useful for materials with heterogeneous distributions and poorly sorted grains with
uniformity coefficient between 1 and 20, and effective grain size between 0.06mm and 0.6mm.

4 g  500  2
k  6  10   lo g    d 10 (1-21)
l  U 

The Slitcher formula is most applicable for grain-sizes between 0.01 mm and 5 mm.

2 g 3 .2 8 7 2
k  1 10  n  d 10 (1-22)
l

The Terzaghi (1964) formula is most applicable for coarse sand. The Terzaghi equation:

2
g  n  0 .1 3  2
k  Ct     d 10 (1-23)
l  3 1  n 

3 3
Where the Ct = sorting coefficient and 6 .1  1 0  C t  1 0 .7  1 0 .

Copyright © Dr.ir. S.A. Miedema TOC Page 25 of 304


Introduction Dredging Engineering.

1.4.5. The Angle of Internal Friction.


Angle of internal friction for a given soil is the angle on the graph (Mohr's Circle) of the shear stress and normal
effective stresses at which shear failure occurs. Angle of Internal Friction, φ, can be determined in the laboratory
by the Direct Shear Test or the Triaxial Stress Test. Typical relationships for estimating the angle of internal
friction, φ, are as follows:

Table 1-7: Empirical values for φ, of granular soils based on the standard penetration number, (From
Bowels, Foundation Analysis).

SPT Penetration, N-Value (blows/ foot) φ (degrees)


0 25 - 30
4 27 - 32
10 30 - 35
30 35 - 40
50 38 - 43

Table 1-8: Relationship between φ, and standard penetration number for sands,
(From Peck 1974, Foundation Engineering Handbook).

SPT Penetration, N-Value (blows/ foot) Density of Sand φ (degrees)


<4 Very loose <29
4 - 10 Loose 29 - 30
10 - 30 Medium 30 - 36
30 - 50 Dense 36 - 41
>50 Very dense >41

Table 1-9: Relationship between φ, and standard penetration number for sands,
(From Meyerhof 1956, Foundation Engineering Handbook).

SPT Penetration, N-Value (blows/ foot) Density of Sand φ (degrees)


<4 Very loose <30
4 - 10 Loose 30 - 35
10 - 30 Medium 35 - 40
30 - 50 Dense 40 - 45
>50 Very dense >45

Lambe & Whitman (1979), page 148 (Figure 1-23) give the relation between the SPT value and the angle of
internal friction, also in a graph. This graph is valid up to 12 m in dry soil. With respect to the internal friction, the
relation given in the graph has an accuracy of 3 degrees. A load of 12 m dry soil with a density of 1.67 ton/m 3
equals a hydrostatic pressure of 20 m.w.c. An absolute hydrostatic pressure of 20 m.w.c. equals 10 m of water
depth if cavitation is considered. Measured SPT values at any depth will have to be reduced to the value that would
occur at 10 m water depth. This can be accomplished with the following equation:

1
S P T1 0   S P Tz (1-24)
 0 .6 4 6  0 .0 3 5 4  z 

Page 26 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Basic Soil Mechanics.

Friction angle versus SPT value.


46

44
Angle of internal friction in degrees

42

40

38

36

34

32

30

28

26
0 10 20 30 40 50 60 70 80 90 100
SPT value in blows/305 mm
Original Fitting Fitting+3 Fitting-3

Figure 1-23: Friction angle versus SPT value (Miedema (1995).

With the aim of curve-fitting, the relation between the SPT value reduced to 10 m water depth and the angle of
internal friction can be summarized to:

 0 .0 1 7 5 3  S P T 1 0
  5 1 .5  2 5 .9  e (1-25)

1.4.6. The Angle of External Friction.

The external friction angle, , or friction between a soil medium and a material such as the composition from a
retaining wall or pile may be expressed in degrees as the following:

Table 1-10: External friction angle φ values.

20º steel piles (NAVFAC)


0.67·φ-0.83·φ USACE
20º steel (Broms)
3/4·φ concrete (Broms)
2/3·φ timber (Broms)
2/3·φ Lindeburg
2/3·φ for concrete walls (Coulomb)

The external friction angle can be estimated as 1/3·φ for smooth retaining walls like sheet piles or concrete surfaces
against timber formwork, or as 1/2·φ to 2/3·φ for rough surfaces. In the absence of detailed information the
assumption of 2/3·φ is commonly made.

Copyright © Dr.ir. S.A. Miedema TOC Page 27 of 304


Introduction Dredging Engineering.

1.4.7. Shear Strength.

1.4.7.1. Introduction.
Shear strength is a term used in soil mechanics to describe the magnitude of the shear stress that a soil can sustain.
The shear resistance of soil is a result of friction and interlocking of particles, and possibly cementation or bonding
at particle contacts. Due to interlocking, particulate material may expand or contract in volume as it is subject to
shear strains. If soil expands its volume, the density of particles will decrease and the strength will decrease; in
this case, the peak strength would be followed by a reduction of shear stress. The stress-strain relationship levels
off when the material stops expanding or contracting, and when inter-particle bonds are broken. The theoretical
state at which the shear stress and density remain constant while the shear strain increases may be called the critical
state, steady state, or residual strength.
The volume change behavior and inter-particle friction depend on the density of the particles, the inter-granular
contact forces, and to a somewhat lesser extent, other factors such as the rate of shearing and the direction of the
shear stress. The average normal inter-granular contact force per unit area is called the effective stress.
If water is not allowed to flow in or out of the soil, the stress path is called an undrained stress path. During
undrained shear, if the particles are surrounded by a nearly incompressible fluid such as water, then the density of
the particles cannot change without drainage, but the water pressure and effective stress will change. On the other
hand, if the fluids are allowed to freely drain out of the pores, then the pore pressures will remain constant and the
test path is called a drained stress path. The soil is free to dilate or contract during shear if the soil is drained. In
reality, soil is partially drained, somewhere between the perfectly undrained and drained idealized conditions. The
shear strength of soil depends on the effective stress, the drainage conditions, the density of the particles, the rate
of strain, and the direction of the strain.
For undrained, constant volume shearing, the Tresca theory may be used to predict the shear strength, but for
drained conditions, the Mohr–Coulomb theory may be used.
Two important theories of soil shear are the critical state theory and the steady state theory. There are key
differences between the steady state condition and the steady state condition and the resulting theory corresponding
to each of these conditions.

1.4.7.2. Undrained Shear Strength.


This term describes a type of shear strength in soil mechanics as distinct from drained strength. Conceptually, there
is no such thing as the undrained strength of a soil. It depends on a number of factors, the main ones being:
 Orientation of stresses
 Stress path
 Rate of shearing
 Volume of material (like for fissured clays or rock mass)
Undrained strength is typically defined by Tresca theory, based on Mohr's circle as:

 1   3  2  S u  U .C .S . (1-26)

It is commonly adopted in limit equilibrium analyses where the rate of loading is very much greater than the rate
at which pore water pressures that are generated due to the action of shearing the soil may dissipate. An example
of this is rapid loading of sands during an earthquake, or the failure of a clay slope during heavy rain, and applies
to most failures that occur during construction. As an implication of undrained condition, no elastic volumetric
strains occur, and thus Poisson's ratio is assumed to remain 0.5 throughout shearing. The Tresca soil model also
assumes no plastic volumetric strains occur. This is of significance in more advanced analyses such as in finite
element analysis. In these advanced analysis methods, soil models other than Tresca may be used to model the
undrained condition including Mohr-Coulomb and critical state soil models such as the modified Cam-clay model,
provided Poisson's ratio is maintained at 0.5.

1.4.7.3. Drained Shear Strength.


The drained shear strength is the shear strength of the soil when pore fluid pressures, generated during the course
of shearing the soil, are able to dissipate during shearing. It also applies where no pore water exists in the soil (the
soil is dry) and hence pore fluid pressures are negligible. It is commonly approximated using the Mohr-Coulomb
equation. (It was called "Coulomb's equation" by Karl von Terzaghi in 1942.) combined it with the principle of
effective stress. In terms of effective stresses, the shear strength is often approximated by:

Page 28 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Basic Soil Mechanics.

  c    ta n    (1-27)

The coefficient of friction μ is equal to tan(φ). Different values of friction angle can be defined, including the peak
friction angle, φ'p, the critical state friction angle, φ'cv, or residual friction angle, φ'r.
c’ is called cohesion, however, it usually arises as a consequence of forcing a straight line to fit through measured
values of (τ,σ')even though the data actually falls on a curve. The intercept of the straight line on the shear stress
axis is called the cohesion. It is well known that the resulting intercept depends on the range of stresses considered:
it is not a fundamental soil property. The curvature (nonlinearity) of the failure envelope occurs because the
dilatancy of closely packed soil particles depends on confining pressure.

1.4.7.4. Cohesion (Internal Shear Strength).


Cohesion (in Latin cohaerere "stick or stay together") or cohesive attraction or cohesive force is the action or
property of like molecules sticking together, being mutually attractive. This is an intrinsic property of a substance
that is caused by the shape and structure of its molecules which makes the distribution of orbiting electrons
irregular when molecules get close to one another, creating electrical attraction that can maintain a macroscopic
structure such as a water drop. Cohesive soils are clay type soils. Cohesion is the force that holds together
molecules or like particles within a soil. Cohesion, c, is usually determined in the laboratory from the Direct Shear
Test. Unconfined Compressive Strength, UCS, can be determined in the laboratory using the Triaxial Test or the
Unconfined Compressive Strength Test. There are also correlations for UCS with shear strength as estimated from
the field using Vane Shear Tests. With a conversion of 1 kips/ft2=47.88 kN/m2.

U .C .S .
c  (1-28)
2

Table 1-11: Guide for Consistency of Fine-Grained Soil, NAVFAC 7.02


SPT Penetration (blows/ foot) Estimated Consistency UCS(kPa)
<2 Very Soft <24
2-4 Soft 24 - 48
4-8 Medium 48 - 96
8 - 15 Stiff 96 – 192
15 - 30 Very Stiff 192 – 388
>30 Hard >388

Table 1-12: Empirical Values for Consistency of Cohesive Soil, (from Foundation Analysis, Bowels)
SPT Penetration (blows/ foot) Estimated Consistency UCS (kips/ft2)
0-2 Very Soft 0 - 0.5
2-4 Soft 0.5 - 1.0
4-8 Medium 1.0 - 2.0
8 - 16 Stiff 2.0 - 4.0
16 - 32 Very Stiff 4.0 - 8.0
>32 Hard >8

1.4.7.5. Adhesion (External Shear Strength).


Adhesion is any attraction process between dissimilar molecular species that can potentially bring them in close
contact. By contrast, cohesion takes place between similar molecules.
Adhesion is the tendency of dissimilar particles and/or surfaces to cling to one another (cohesion refers to the
tendency of similar or identical particles/surfaces to cling to one another). The forces that cause adhesion and
cohesion can be divided into several types. The intermolecular forces responsible for the function of various kinds
of stickers and sticky tape fall into the categories of chemical adhesion, dispersive adhesion, and diffusive
adhesion.

Copyright © Dr.ir. S.A. Miedema TOC Page 29 of 304


Introduction Dredging Engineering.

1.4.8. UCS or Unconfined Compressive Strength.


UCS is one of the most basic parameters of rock strength, and the most common determination performed for bore
ability predictions. It is measured in accordance with the procedures given in ASTM D2938, with the length to
diameter ratio of 2 by using NX-size core samples. 3 to 5 UCS determinations are recommended to achieve
statistical significance of the results. If the sample length to diameter ratio was greater or less than 2, ASTM
recommends a correction factor that is applied to the UCS value determined from testing. UCS measurements are
made using an electronic-servo controlled MTS stiff testing machine with a capacity of 220 kips. Loading data
and other test parameters are recorded with a computer based data acquisition system, and the data is subsequently
reduced and analyzed with a customized spreadsheet program.

The most important test for rock in the field of dredging is the uniaxial unconfined compressive strength (UCS).
In the test a cylindrical rock sample is axial loaded till failure. Except the force needed, the deformation is measured
too. So the complete stress-strain curve is measured from which the deformation modulus and the specific work
of failure can be calculated. The unconfined compressive strength of the specimen is calculated by dividing the
maximum load at failure by the sample cross-sectional area:

F
c  (1-29)
A

Figure 1-24: A UCS test facility (Timely Engineering Soil Tests, LLC).

Figure 1-25: Bending (Vlasblom (2003-2007)).

Page 30 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Basic Soil Mechanics.

1.4.9. Unconfined Tensile Strength.


The uniaxial unconfined tensile strength is defined in the same way as the compressive strength. Sample
preparation and testing procedure require much effort and not commonly done. Another method to determine the
tensile strength, also commonly not used, is by bending a sample.

1.4.10. BTS or Brazilian Tensile Strength.


Indirect, or Brazilian, tensile strength is measured using NX-size core samples cut to an approximate 0.5 length-
to-diameter ratio, and following the procedures of ASTM D3967. BTS measurements are made using an
electronic-servo controlled MTS stiff testing machine with a capacity of 220 kips. Loading data and other test
parameters are recorded with a computer based data acquisition system, and the data is subsequently reduced and
analyzed with a customized spreadsheet program. BTS provides a measure of rock toughness, as well as strength.
The indirect tensile strength is calculated as follows (Fairhurst (1964)):

2F
T  (1-30)
 L D

In bedded/foliated rocks, particular attention needs to be given to loading direction with respect to
bedding/foliation. The rock should be loaded so that breakage occurs in approximately the same direction as
fracture propagation between adjacent cuts on the tunnel face. This is very important assessment in mechanical
excavation by tunnel boring machines. The most common used test to estimate, in an indirect way, the tensile
strength is the Brazilian split test. Here the cylindrical sample is tested radial.

The validity of BTS to determine de UTS is discussed by many researchers. In general it can be stated that the
BTS over estimates the UTS. According to Pells (1993) this discussion is in most applications in practice largely
academic.

Figure 1-26: The Brazilian split test (Vlasblom (2003-2007)).

1.4.11. Hardness.
Hardness is a loosely defined term, referring the resistance to rock or minerals against an attacking tool. Hardness
is determined using rebound tests (f.i. Schmidt hammer), indentation tests, (Brinell, Rockwell) or scratch tests
(Mohs). The last test is based on the fact that a mineral higher in the scale can scratch a mineral lower in the scale.

Copyright © Dr.ir. S.A. Miedema TOC Page 31 of 304


Introduction Dredging Engineering.

Although this scale was established in the early of the 19th century it appeared that the increment of Mohs scale
corresponded with a 60% increase in indentation hardness.

Table 1-13: The Mohs scale (source Wikipedia).

Mohs hardness Mineral Chemical formula Absolute hardness. Image

1 Talc Mg3Si4O10(OH)2 1

2 Gypsum CaSO4·2H2O 3

3 Calcite CaCO3 9

4 Fluorite CaF2 21

5 Apatite Ca5(PO4)3(OH−,Cl−,F−) 48

6 Orthoclase KAlSi3O8 72

7 Quartz SiO2 100

8 Topaz Al2SiO4(OH−,F−)2 200

9 Corundum Al2O3 400

10 Diamond C 1600

Page 32 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Basic Soil Mechanics.

1.5. Criteria & Concepts.


1.5.1. Failure Criteria.
After a soil reaches the critical state, it is no longer contracting or dilating and the shear stress on the failure plane
τcrit is determined by the effective normal stress on the failure plane σn' and critical state friction angle, φcv, :

 c r it   n ' ta n   c v  (1-31)

The peak strength of the soil may be greater, however, due to the interlocking (dilatancy) contribution. This may
be stated:


 p e a k   n ' t a n  p e a k  (1-32)

Where φpeak > φcv. However, use of a friction angle greater than the critical state value for design requires care.
The peak strength will not be mobilized everywhere at the same time in a practical problem such as a foundation,
slope or retaining wall. The critical state friction angle is not nearly as variable as the peak friction angle and hence
it can be relied upon with confidence. Not recognizing the significance of dilatancy, Coulomb proposed that the
shear strength of soil may be expressed as a combination of adhesion and friction components:

   ' ta n     c ' (1-33)

It is now known that the c' and φ parameters in the last equation are not fundamental soil properties. In particular,
c' and φ are different depending on the magnitude of effective stress. According to Schofield (2006), the
longstanding use of c' in practice has led many engineers to wrongly believe that c' is a fundamental parameter.
This assumption that c' and φ are constant can lead to overestimation of peak strengths.

1.5.2. The Phi=0 Concept.


When a fast triaxial test is carried out, so an unconsolidated undrained test, it is very well possible that the pore
pressures will be equal to the increase of the cell pressure. If a test at high cell pressure is carried out, the only
difference with a test with a low cell pressure is the value of the pore pressures. The grain pressures will be almost
equal in both cases and the result is, that we will find the same critical Mohr circle. So let’s consider a series of
unconsolidated undrained (UU) tests. Three specimens are selected and all are consolidated to 110 kPa. This brings
the specimens to the end of step 1 in the UU test program. Now the confining pressures are changed to say 70, 140
and 700 kPa, without allowing further consolidation and the sheared undrained. The result, within experimental
scatter, is that the shear stress or radius of the Mohr circle is about 35 kPa for each specimen.
So what happened?
When the confining pressure was changed, the pore pressure in the fully saturated specimens changed just as much
as did the confining pressure, and the effective stress remained unchanged and equal in each specimen. Thus the
effective stress remained 110 kPa and each specimen behaved during shear just as did the CU specimen. The shear
stress and thus the radius of the Mohr circle did not increase and apparently the specimens did not encounter
internal friction. This is called the phi=0 concept. In clays with a very low permeability and at a high deformation
rate, like during the cutting of clay, the clay behaves like the internal friction angle is zero. So for cutting processes
the phi=0 concept will be applied.

1.5.3. Factors Controlling Shear Strength of Soils.


The stress-strain relationship of soils, and therefore the shearing strength, is affected by:
1. Soil composition (basic soil material): mineralogy, grain size and grain size distribution, shape of
particles, pore fluid type and content, ions on grain and in pore fluid.
2. State (initial): Defined by the initial void ratio, effective normal stress and shear stress (stress history).
State can be described by terms such as: loose, dense, over consolidated, normally consolidated, stiff,
soft, contractive, dilative, etc.
3. Structure: Refers to the arrangement of particles within the soil mass; the manner the particles are packed
or distributed. Features such as layers, joints, fissures, slickensides, voids, pockets, cementation, etc., are
part of the structure. Structure of soils is described by terms such as: undisturbed, disturbed, remolded,

Copyright © Dr.ir. S.A. Miedema TOC Page 33 of 304


Introduction Dredging Engineering.

compacted, cemented; flocculent, honey-combed, single-grained; flocculated, deflocculated; stratified,


layered, laminated; isotropic and anisotropic.
4. Loading conditions: Effective stress path, i.e., drained, and undrained; and type of loading, i.e.,
magnitude, rate (static, dynamic), and time history (monotonic, cyclic).

The shear strength and stiffness of soil determines whether or not soil will be stable or how much it will deform.
Knowledge of the strength is necessary to determine if a slope will be stable, if a building or bridge might settle
too far into the ground, and the limiting pressures on a retaining wall. It is important to distinguish between failure
of a soil element and the failure of a geotechnical structure (e.g., a building foundation, slope or retaining wall);
some soil elements may reach their peak strength prior to failure of the structure. Different criteria can be used to
define the "shear strength" and the "yield point" for a soil element from a stress-strain curve. One may define the
peak shear strength as the peak of a stress strain curve, or the shear strength at critical state as the value after large
strains when the shear resistance levels off. If the stress-strain curve does not stabilize before the end of shear
strength test, the "strength" is sometimes considered to be the shear resistance at 15% to 20% strain. The shear
strength of soil depends on many factors including the effective stress and the void ratio.
The shear stiffness is important, for example, for evaluation of the magnitude of deformations of foundations and
slopes prior to failure and because it is related to the shear wave velocity. The slope of the initial, nearly linear,
portion of a plot of shear stress as a function of shear strain is called the shear modulus

1.5.4. Friction, Interlocking & Dilation.


Soil is an assemblage of particles that have little to no cementation while rock (such as sandstone) may consist of
an assembly of particles that are strongly cemented together by chemical bonds. The shear strength of soil is
primarily due to inter-particle friction and therefore, the shear resistance on a plane is approximately proportional
to the effective normal stress on that plane.[3] But soil also derives significant shear resistance from interlocking of
grains. If the grains are densely packed, the grains tend to spread apart from each other as they are subject to shear
strain. The expansion of the particle matrix due to shearing was called dilatancy by Osborne Reynolds. If one
considers the energy required to shear an assembly of particles there is energy input by the shear force, T, moving
a distance, x and there is also energy input by the normal force, N, as the sample expands a distance, y. Due to the
extra energy required for the particles to dilate against the confining pressures, dilatant soils have greater peak
strength than contractive soils. Furthermore, as dilative soil grains dilate, they become looser (their void ratio
increases), and their rate of dilation decreases until they reach a critical void ratio. Contractive soils become denser
as they shear, and their rate of contraction decreases until they reach a critical void ratio.
The tendency for a soil to dilate or contract depends primarily on the confining pressure and the void ratio of the
soil. The rate of dilation is high if the confining pressure is small and the void ratio is small. The rate of contraction
is high if the confining pressure is large and the void ratio is large. As a first approximation, the regions of
contraction and dilation are separated by the critical state line.

1.5.5. Effective Stress.


Karl von Terzaghi (1964) first proposed the relationship for effective stress in 1936. For him, the term ‘effective’
meant the calculated stress that was effective in moving soil, or causing displacements. It represents the average
stress carried by the soil skeleton. Effective stress (σ') acting on a soil is calculated from two parameters, total
stress (σ) and pore water pressure (u) according to:

'  u (1-34)

Typically, for simple examples:

   s o il  H s o il a n d u =  w  H w (1-35)

Much like the concept of stress itself, the formula is a construct, for the easier visualization of forces acting on a
soil mass, especially simple analysis models for slope stability, involving a slip plane. With these models, it is
important to know the total weight of the soil above (including water), and the pore water pressure within the slip
plane, assuming it is acting as a confined layer.
However, the formula becomes confusing when considering the true behavior of the soil particles under different
measurable conditions, since none of the parameters are actually independent actors on the particles.

Consider a grouping of round quartz sand grains, piled loosely, in a classic ‘cannonball’ arrangement. As can be
seen, there is a contact stress where the spheres actually touch. Pile on more spheres and the contact stresses

Page 34 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Basic Soil Mechanics.

increase, to the point of causing frictional instability (dynamic friction), and perhaps failure. The independent
parameter affecting the contacts (both normal and shear) is the force of the spheres above. This can be calculated
by using the overall average density of the spheres and the height of spheres above.
If we then have these spheres in a beaker and add some water, they will begin to float a little depending on their
density (buoyancy). With natural soil materials, the effect can be significant, as anyone who has lifted a large rock
out of a lake can attest. The contact stress on the spheres decreases as the beaker is filled to the top of the spheres,
but then nothing changes if more water is added. Although the water pressure between the spheres (pore water
pressure) is increasing, the effective stress remains the same, because the concept of 'total stress' includes the
weight of all the water above. This is where the equation can become confusing, and the effective stress can be
calculated using the buoyant density of the spheres (soil), and the height of the soil above.

The concept of effective stress truly becomes interesting when dealing with non-hydrostatic pore water pressure.
Under the conditions of a pore pressure gradient, the ground water flows, according to the permeability equation
(Darcy's law). Using our spheres as a model, this is the same as injecting (or withdrawing) water between the
spheres. If water is being injected, the seepage force acts to separate the spheres and reduces the effective stress.
Thus, the soil mass becomes weaker. If water is being withdrawn, the spheres are forced together and the effective
stress increases. Two extremes of this effect are quicksand, where the groundwater gradient and seepage force act
against gravity; and the 'sandcastle effect', where the water drainage and capillary action act to strengthen the sand.
As well, effective stress plays an important role in slope stability, and other geotechnical engineering and
engineering geology problems, such as groundwater-related subsidence.

1.5.6. Pore Water Pressure: Hydrostatic Conditions.


If there is no pore water flow occurring in the soil, the pore water pressures will be hydrostatic. The water table is
located at the depth where the water pressure is equal to the atmospheric pressure. For hydrostatic conditions, the
water pressure increases linearly with depth below the water table:

u  w  g  zw (1-36)

1.5.7. Pore Water Pressure: Capillary Action.


Due to surface tension water will rise up in a small capillary tube above a free surface of water. Likewise, water
will rise up above the water table into the small pore spaces around the soil particles. In fact the soil may be
completely saturated for some distance above the water table. Above the height of capillary saturation, the soil
may be wet but the water content will decrease with elevation. If the water in the capillary zone is not moving, the
water pressure obeys the equation of hydrostatic equilibrium, u = ρw·g·zw, but note that zw, is negative above the
water table. Hence, hydrostatic water pressures are negative above the water table. The thickness of the zone of
capillary saturation depends on the pore size, but typically, the heights vary between a centimeter or so for coarse
sand to tens of meters for a silt or clay.
The surface tension of water explains why the water does not drain out of a wet sand castle or a moist ball of clay.
Negative water pressures make the water stick to the particles and pull the particles to each other, friction at the
particle contacts make a sand castle stable. But as soon as a wet sand castle is submerged below a free water
surface, the negative pressures are lost and the castle collapses. Considering the effective stress equation, σ' = σ −
u,, if the water pressure is negative, the effective stress may be positive, even on a free surface (a surface where
the total normal stress is zero). The negative pore pressure pulls the particles together and causes compressive
particle to particle contact forces.
Negative pore pressures in clayey soil can be much more powerful than those in sand. Negative pore pressures
explain why clay soils shrink when they dry and swell as they are wetted. The swelling and shrinkage can cause
major distress, especially to light structures and roads.

1.5.8. Darcy’s Law.


Darcy's law states that the volume of flow of the pore fluid through a porous medium per unit time is proportional
to the rate of change of excess fluid pressure with distance. The constant of proportionality includes the viscosity
of the fluid and the intrinsic permeability of the soil.

K  A ub  ua 
Q   (1-37)
l L

Copyright © Dr.ir. S.A. Miedema TOC Page 35 of 304


Introduction Dredging Engineering.

The negative sign is needed because fluids flow from high pressure to low pressure. So if the change in pressure
is negative (in the x-direction) then the flow will be positive (in the x-direction). The above equation works well
for a horizontal tube, but if the tube was inclined so that point b was a different elevation than point a, the equation
would not work. The effect of elevation is accounted for by replacing the pore pressure by excess pore pressure,
ue defined as:

uc  u  w  g  z (1-38)

Where z is the depth measured from an arbitrary elevation reference (datum). Replacing u by ue we obtain a more
general equation for flow:

K  A  u c ,b  u c ,a 
Q   (1-39)
l L

Figure 1-27: Diagram showing definitions and directions for Darcy’s law.

Dividing both sides of the equation by A, and expressing the rate of change of excess pore pressure as a derivative,
we obtain a more general equation for the apparent velocity in the x-direction:

K duc
qx   (1-40)
l dx

Where qx has units of velocity and is called the Darcy velocity, or discharge velocity. The seepage velocity (vsx =
average velocity of fluid molecules in the pores) is related to the Darcy velocity, and the porosity, n:

qx
v s ,x  (1-41)
n

Civil engineers predominantly work on problems that involve water and predominantly work on problems on earth
(in earth’s gravity). For this class of problems, civil engineers will often write Darcy's law in a much simpler form:

q x  k  ix (1-42)

Where k is called permeability, and is defined as:

l  g
k  K  (1-43)
l
And i is called the hydraulic gradient. The hydraulic gradient is the rate of change of total head with distance.
Values are for typical fresh groundwater conditions, using standard values of viscosity and specific gravity for
water at 20°C and 1 atm.

Page 36 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Basic Soil Mechanics.

Table 1-14: Typical values of the permeability k.

Soil Permeability (m/s) Degree of permeability


Well sorted gravel 100>k>10-2 Extremely high
Gravel 10-2>k>10-3 Very high
Sandy gravel, clean sand, 10-3>k>10-5 High to Medium
fine sand
Sand, dirty sand, silty 10-5>k>10-7 Low
sand
Silt, silty clay 10-7>k>10-9 Very low
Clay <10-9 Vitually impermeable
Highly fractured rocks 10 >k>10-3
0
Very high
Oil reservoir rocks 10-4>k>10-6 Medium to Low
Fresh sandstone 10-7>k>10-8 Very low
Fresh limestone, dolomite 10-9>k>10-10 Vitually impermeable
Fresh granite <10-11 Vitually impermeable

Table 1-15: Some permeabilities according to Hazen’s equation.

Material Permeability (m/s) d10 (mm)


Uniform coarse sand 0.0036 0.6
Uniform medium sand 0.0009 0.3
Clean, well-graded sand 0.0001 0.1
Uniform fine sand 36·10-6 0.06
Well-graded fine sand 4·10-6 0.02
Silty sand 10-6 0.01
Uniform silt 36·10-8 0.006
Sandy clay 4·10-8 0.002
Silty clay 10-8 0.001
Clay 64·10-10 0.0008
Colloidal clay 9·10-11 0.00003

1.5.9. Brittle versus Ductile Failure.


The terms ductile failure and brittle failure are often used in literature for the failure of materials with shear strength
and tensile strength.

“In materials science, ductility is a solid material's ability to deform under tensile stress; this is often characterized
by the material's ability to be stretched into a wire. Malleability, a similar property, is a material's ability to deform
under compressive stress; this is often characterized by the material's ability to form a thin sheet by hammering
or rolling. Both of these mechanical properties are aspects of plasticity, the extent to which a solid material can
be plastically deformed without fracture. Ductility and malleability are not always coextensive – for instance,
while gold has high ductility and malleability, lead has low ductility but high malleability. The word ductility is
sometimes used to embrace both types of plasticity.
A material is brittle if, when subjected to stress, it breaks without significant deformation (strain). Brittle materials
absorb relatively little energy prior to fracture, even those of high strength. Breaking is often accompanied by a
snapping sound. Brittle materials include most ceramics and glasses (which do not deform plastically) and some
polymers, such as PMMA and polystyrene. Many steels become brittle at low temperatures (see ductile-brittle
transition temperature), depending on their composition and processing. When used in materials science, it is
generally applied to materials that fail when there is little or no evidence of plastic deformation before failure.
One proof is to match the broken halves, which should fit exactly since no plastic deformation has occurred.
Generally, the brittle strength of a material can be increased by pressure. This happens as an example in the
brittle-ductile transition zone at an approximate depth of 10 kilometers in the Earth's crust, at which rock becomes
less likely to fracture, and more likely to deform ductile.” (Source Wikipedia).
In rock failure a distinction is made between brittle, brittle ductile and ductile failure. Factors determining those
types of failure are the ductility number (ratio compressive strength over tensile strength), the confining pressure
and the temperature. During dredging the temperature will have hardly any influence, however when drilling deep
oil wells temperature will play an important role. The confining pressure, where the failure transit from brittle to
ductile is called bp.

Copyright © Dr.ir. S.A. Miedema TOC Page 37 of 304


Introduction Dredging Engineering.

Figure 1-28: Brittle failure types (Vlasblom (2003-2007)).

Brittle failure occurs at relative low confining pressures 3 < bp en deviator stress q=1-3 > ½qu. The strength
increases with the confining pressure, but decreases after the peak strength to a residual value. The presence of
pore water can play an important role.
Brittle failure types are:
 Pure tensile failure with or without a small confining pressure.
 Axial tensile failure
 Shear plane failure
Brittle ductile failure is also called semi brittle. In the transition area where 3  bp, the deformations are not
restricted to local shear planes or fractures but are divided over the whole area. The residual- strength is more or
less equal to the peak strength.
Ductile failure. A rock fails ductile when 3 >> qu and 3 > bp while the force stays constant or increases some
what with increasing deformation.

Figure 1-29: Brittle-ductile failure of marble (M.S. Patterson, Australian National University).

Page 38 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Basic Soil Mechanics.

1.6. Soil Mechanical Tests.

1.6.1. Sieve Analysis.


The size distribution of gravel and sand particles are typically measured using sieve analysis. The formal procedure
is described in ASTM D6913-04(2009). A stack of sieves with accurately dimensioned holes between a mesh of
wires is used to separate the particles into size bins. A known volume of dried soil, with clods broken down to
individual particles, is put into the top of a stack of sieves arranged from coarse to fine. The stack of sieves is
shaken for a standard period of time so that the particles are sorted into size bins. This method works reasonably
well for particles in the sand and gravel size range. Fine particles tend to stick to each other, and hence the sieving
process is not an effective method. If there are a lot of fines (silt and clay) present in the soil it may be necessary
to run water through the sieves to wash the coarse particles and clods through.
A variety of sieve sizes are available. The boundary between sand and silt is arbitrary. According to the Unified
Soil Classification System, a #4 sieve (4 openings per inch) having 4.75mm opening size separates sand from
gravel and a #200 sieve with an 0.075 mm opening separates sand from silt and clay. According to the British
standard, 0.063 mm is the boundary between sand and silt, and 2 mm is the boundary between sand and gravel.

Figure 1-30: A set of sieves (Essa Australia from: www.directindustry.com).

1.6.2. Hydrometer Analysis.


The classification of fine-grained soils, i.e., soils that are finer than sand, is determined primarily by their Atterberg
limits, not by their grain size. If it is important to determine the grain size distribution of fine-grained soils, the
hydrometer test may be performed. In the hydrometer tests, the soil particles are mixed with water and shaken to
produce a dilute suspension in a glass cylinder, and then the cylinder is left to sit. A hydrometer is used to measure
the density of the suspension as a function of time. Clay particles may take several hours to settle past the depth
of measurement of the hydrometer. Sand particles may take less than a second. Stoke's law provides the theoretical
basis to calculate the relationship between sedimentation velocity and particle size. ASTM provides the detailed
procedures for performing the Hydrometer test.
Clay particles can be sufficiently small that they never settle because they are kept in suspension by Brownian
motion, in which case they may be classified as colloids.

Copyright © Dr.ir. S.A. Miedema TOC Page 39 of 304


Introduction Dredging Engineering.

1.6.3. Standard Penetration Test.


The standard penetration test (SPT) is an in-situ dynamic penetration test designed to provide information on the
geotechnical engineering properties of soil. The test procedure is described in the British Standard BS EN ISO
22476-3, ASTM D1586 and Australian Standards AS 1289.6.3.1.

Figure 1-31: The Standard Penetration Test (www.shalviengineering.com).

The test uses a thick-walled sample tube, with an outside diameter of 50 mm and an inside diameter of 35 mm,
and a length of around 650 mm. This is driven into the ground at the bottom of a borehole by blows from a slide
hammer with a weight of 63.5 kg (140 lb) falling through a distance of 760 mm (30 in). The sample tube is driven
150 mm into the ground and then the number of blows needed for the tube to penetrate each 150 mm (6 in) up to
a depth of 450 mm (18 in) is recorded. The sum of the number of blows required for the second and third 6 in. of
penetration is termed the "standard penetration resistance" or the "N-value". In cases where 50 blows are
insufficient to advance it through a 150 mm (6 in) interval the penetration after 50 blows is recorded. The blow
count provides an indication of the density of the ground, and it is used in many empirical geotechnical engineering
formulae.

The main purpose of the test is to provide an indication of the relative density of granular deposits, such as sands
and gravels from which it is virtually impossible to obtain undisturbed samples. The great merit of the test, and
the main reason for its widespread use is that it is simple and inexpensive. The soil strength parameters which can
be inferred are approximate, but may give a useful guide in ground conditions where it may not be possible to
obtain borehole samples of adequate quality like gravels, sands, silts, clay containing sand or gravel and weak
rock. In conditions where the quality of the undisturbed sample is suspect, e.g. very silty or very sandy clays, or
hard clays, it is often advantageous to alternate the sampling with standard penetration tests to check the strength.
If the samples are found to be unacceptably disturbed, it may be necessary to use a different method for measuring
strength like the plate test. When the test is carried out in granular soils below groundwater level, the soil may
become loosened. In certain circumstances, it can be useful to continue driving the sampler beyond the distance
specified, adding further drilling rods as necessary. Although this is not a standard penetration test, and should not

Page 40 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Basic Soil Mechanics.

be regarded as such, it may at least give an indication as to whether the deposit is really as loose as the standard
test may indicate.
The usefulness of SPT results depends on the soil type, with fine-grained sands giving the most useful results, with
coarser sands and silty sands giving reasonably useful results, and clays and gravelly soils yielding results which
may be very poorly representative of the true soil conditions. Soils in arid areas, such as the Western United States,
may exhibit natural cementation. This condition will often increase the standard penetration value.
The SPT is used to provide results for empirical determination of a sand layer's susceptibility to earthquake
liquefaction, based on research performed by Harry Seed, T. Leslie Youd, and others.
Despite its many flaws, it is usual practice to correlate SPT results with soil properties relevant for geotechnical
engineering design. The reason being that SPT results are often the only test results available, therefore the use of
direct correlations has become common practice in many countries.
Different correlations are proposed for granular and cohesive soils.

1.6.4. Cone Penetration Test.


The cone penetration test (CPT) is an in situ testing method used to determine the geotechnical engineering
properties of soils and delineating soil stratigraphy. It was initially developed in the 1950s at the Dutch Laboratory
for Soil Mechanics in Delft to investigate soft soils. Based on this history it has also been called the "Dutch cone
test". Today, the CPT is one of the most used and accepted in situ test methods for soil investigation worldwide.
The test method consists of pushing an instrumented cone, with the tip facing down, into the ground at a controlled
rate (usually 2 centimeters/second). The resolution of the CPT in delineating stratigraphic layers is related to the
size of the cone tip, with typical cone tips having a cross-sectional area of either 10 or 15 cm², corresponding to
diameters of 3.6 and 4.4 cm.
The early applications of CPT mainly determined the soil geotechnical property of bearing capacity. The original
cone penetrometers involved simple mechanical measurements of the total penetration resistance to pushing a tool
with a conical tip into the soil. Different methods were employed to separate the total measured resistance into
components generated by the conical tip (the "tip friction") and friction generated by the rod string. A friction
sleeve was added to quantify this component of the friction and aid in determining soil cohesive strength in the
1960s (Begemann, 1965). Electronic measurements began in 1948 and improved further in the early 1970s (de
Reister, 1971). Most modern electronic CPT cones now also employ a pressure transducer with a filter to gather
pore water pressure data. The filter is usually located either on the cone tip (the so-called U1 position), immediately
behind the cone tip (the most common U2 position) or behind the friction sleeve (U3 position). Pore water pressure
data aids determining stratigraphy and is primarily used to correct tip friction values for those effects. CPT testing
which also gathers this piezometer data is called CPTU testing. CPT and CPTU testing equipment generally
advances the cone using hydraulic rams mounted on either a heavily ballasted vehicle or using screwed-in anchors
as a counter-force. One advantage of CPT over the Standard Penetration Test (SPT) is a more continuous profile
of soil parameters, with CPTU data recorded typically at 2cm intervals.
In addition to the mechanical and electronic cones, a variety of other CPT-deployed tools have been developed
over the years to provide additional subsurface information. One common tool advanced during CPT testing is a
geophone set to gather seismic shear wave and compression wave velocities. This data helps determine the shear
modulus and Poisson's ratio at intervals through the soil column for soil liquefaction analysis and low-strain soil
strength analysis. Engineers use the shear wave velocity and shear modulus to determine the soil's behavior under
low-strain and vibratory loads. Additional tools such as laser-induced fluorescence, X-ray fluorescence[1], soil
conductivity/resistivity, membrane interface probe and cameras for capturing video imagery are also increasingly
advanced in conjunction with the CPT probe. An additional CPT deployed tool used in Britain, Netherlands,
Germany, Belgium and France is a piezocone combined with a tri-axial magnetometer. This is used to attempt to
ensure that tests, boreholes, and piles, do not encounter unexploded ordnance (UXB) or duds. The magnetometer
in the cone detects ferrous materials of 50 kg or larger within a radius of up to about 2 m distance from the probe
depending on the material, orientation and soil conditions.
CPT for geotechnical applications was standardized in 1986 by ASTM Standard D 3441 (ASTM, 2004). ISSMGE
provides international standards on CPT and CPTU. Later ASTM Standards have addressed the use of CPT for
various environmental site characterization and groundwater monitoring activities. Particularly for geotechnical
soil investigations, CPT is gaining popularity compared to standard penetration testing as a method of geotechnical
soil investigation by its increased accuracy, speed of deployment, more continuous soil profile and reduced cost
over other soil testing methods. The ability to advance additional in situ testing tools using the CPT direct push
drilling rig, including the seismic tools described above, are accelerating this process.

Copyright © Dr.ir. S.A. Miedema TOC Page 41 of 304


Introduction Dredging Engineering.

Figure 1-32: A typical CPT test setup (www.geotechdata.com).

Figure 1-33: Several configurations of cones (www.geotechdata.info).

Page 42 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Basic Soil Mechanics.

Figure 1-34: Several cone configurations.

1.6.5. Triaxial Test.


A triaxial shear test is a common method to measure the mechanical properties of many deformable solids,
especially soil (e.g. sand, clay) and rock, and other granular materials or powders. There are several variations on
the test. Although the name triaxial test suggests that the stresses would be different in three directions, this is not
true in the test as is usually done. In this test with oil or water as confining medium, the confining pressures are
equal in all directions (i.e. in terms of principal stresses: for a compression test: σ1 ≠ σ2 = σ3 and for tensile: σ1 =
σ2 ≠ σ3). Only in a true triaxial test the stresses in all three directions can be different (i.e. σ1 ≠ σ2 ≠ σ3). For loose
granular materials like sand or gravel, the material is contained in a cylindrical latex sleeve with a flat, circular
metal plate or platen closing off the top and bottom ends. This cylinder is placed into a bath of water (mostly water
but may be any other fluid) to provide pressure along the sides of the cylinder. The top platen can then be
mechanically driven up or down along the axis of the cylinder to squeeze the material. The distance that the upper
platen travels is measured as a function of the force required to move it, as the pressure of the surrounding water
is carefully controlled. The net change in volume of the material is also measured by how much water moves in or
out of the surrounding bath. The test for cohesive (non-loose) materials (e.g. clay, rock) is similar to the test for
loose granular materials. For rock testing the sleeve may be a thin metal sheeting rather than latex. Triaxial testing
on rock is fairly seldom done because the high forces and pressures required to break a rock sample imply very
costly and cumbersome testing equipment available at few laboratories in the world. During the test the pore
pressures of fluids (e.g. water, oil) or gasses in the sample may be measured.

The principle behind a triaxial shear test is that the stress applied in the vertical direction (along the axis of the
cylindrical sample) can be different from the stresses applied in the horizontal directions perpendicular to the sides
of the cylinder, i.e. the confining pressure). In a homogeneous and isotropic material this produces a non-
hydrostatic stress state, with shear stress that may lead to failure of the sample in shear. In non-homogeneous and
anisotropic samples (e.g. bedded or jointed samples) failure may occur due to bending moments and, hence, failure
may be tensile. Also combinations of bending and shear failure may happen in inhomogeneous and anisotropic
material.
A solid is defined as a material that can support shear stress without moving. However, every solid has an upper
limit to how much shear stress it can support. The triaxial test is designed to measure that limit. The stress on the
platens is increased until the material in the cylinder fails and forms sliding regions within itself, known as shear
bands. A motion where a material is deformed under shear stress is known as shearing. The geometry of the
shearing in a triaxial test typically causes the sample to become shorter while bulging out along the sides. The

Copyright © Dr.ir. S.A. Miedema TOC Page 43 of 304


Introduction Dredging Engineering.

stress on the platen is then reduced and the water pressure pushes the sides back in, causing the sample to grow
taller again. This cycle is usually repeated several times while collecting stress and strain data about the sample.
During the shearing, a granular material will typically have a net gain or loss of volume. If it had originally been
in a dense state, then it typically gains volume, a characteristic known as Reynolds' dilatancy. If it had originally
been in a very loose state, then contraction may occur before the shearing begins or in conjunction with the
shearing.

Figure 1-35: The Triaxial apparatus (www.geotechdata.info).

From the triaxial test data, it is possible to extract fundamental material parameters about the sample, including its
angle of shearing resistance, apparent cohesion, and dilatancy angle. These parameters are then used in computer
models to predict how the material will behave in a larger-scale engineering application. An example would be to
predict the stability of the soil on a slope, whether the slope will collapse or whether the soil will support the shear
stresses of the slope and remain in place. Triaxial tests are used along with other tests to make such engineering
predictions.
The triaxial test can be used to determine the shear strength of a discontinuity. A homogeneous and isotropic
sample (see above) fails due to shear stresses in the sample. If a sample with a discontinuity is orientated such that
the discontinuity is about parallel to the plane in which maximum shear stress will be developed during the test,
the sample will fail due to shear displacement along the discontinuity, and hence, the shear strength of a
discontinuity can be calculated.

There are several variations of the triaxial test:

Page 44 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Basic Soil Mechanics.

1.6.5.1. Consolidated Drained (CD).


In a consolidated drained test the sample is consolidated and sheared in compression with drainage. The rate of
axial deformation is kept constant, i.e. is strain controlled. The idea is that the test allows the sample and the pore
pressures to fully consolidate (i.e. adjust) to the surrounding stresses. The test may take a long time to allow the
sample to adjust, in particular low permeability samples need a long time to drain and adjust stain to stress levels.

1.6.5.2. Consolidated Undrained (CU).


In a consolidated undrained test the sample is not allowed to drain. The shear characteristics are measured under
undrained conditions and the sample is assumed to be fully consolidated under the stresses applied that should be
similar to the field conditions. Test in particular used if a change in stress is to happen without time for further
consolidation.

1.6.5.3. Unconsolidated Undrained (UU).


In an unconsolidated undrained test the sample is not allowed to drain. The sample is compressed at a constant
rate (strain-controlled).

Figure 1-36: The Triaxial apparatus cross-section (civilblog.org).

1.6.6. Shear Test.


A direct shear test also known as shear box test is a laboratory or field test used by geotechnical engineers to
measure the shear strength properties of soil or rock material, or of discontinuities in soil or rock masses. For soil
the U.S. and U.K. standards defining how the test should be performed are ASTM D 3080 and BS 1377-7:1990
respectively to establish the shear strength properties of soil. It is also possible to estimate typical values of the
shear strength parameters based on the type and classification of the soils. For rock the test is generally restricted
to rock with (very) low (shear) strength. The test is, however, standard practice to establish the shear strength
properties of discontinuities in rock.

Copyright © Dr.ir. S.A. Miedema TOC Page 45 of 304


Introduction Dredging Engineering.

The test is performed on three or four specimens from a relatively undisturbed soil sample. A specimen is placed
in a shear box which has two stacked rings to hold the sample; the contact between the two rings is at approximately
the mid-height of the sample. A confining stress is applied vertically to the specimen, and the upper ring is pulled
laterally until the sample fails, or through a specified strain. The load applied and the strain induced is recorded at
frequent intervals to determine a stress-strain curve for the confining stress.
Direct Shear tests can be performed under several conditions. The sample is normally saturated before the test is
run, but can be run at the in-situ moisture content. The rate of strain can be varied to create a test of undrained or
drained conditions, depending whether the strain is applied slowly enough for water in the sample to prevent pore-
water pressure buildup.

Figure 1-37: The direct shear test.

Several specimens are tested at varying confining stresses to determine the shear strength parameters, the soil
cohesion (c) and the angle of internal friction (commonly friction angle) (φ). The results of the tests on each
specimen are plotted on a graph with the peak (or residual) stress on the x-axis and the confining stress on the y-
axis. The y-intercept of the curve which fits the test results is the cohesion, and the slope of the line or curve is the
friction angle.

1.6.7. Point Load Test.


The Point Load Strength test is intended as an index test for the strength classification of rock materials. It may
also be used to predict other strength parameters with which it is correlated, for example the unconfined
compressive and the tensile strength. It is measured in accordance with the procedures recommended in ASTM
D5731, usually with NX-size core samples. The testing machine consists of a loading frame, which measures the
force required to break the sample, and a system for measuring the distance between the two platen contact points.
Rock specimens in the form of either core, cut blocks, or irregular lumps are broken by application of concentrated
load through a pair of spherically truncated, conical platens. The applied force at failure of the sample and distance
between the platen tips are recorded in order to calculate the point load index as follows:

F
Is  (1-44)
2
De

Another test that is familiar with the Brazilian splitting test is the point load strength test. This test is executed
either axial, diametrical or on irregular pieces. The point load test is frequently used to determine the strength when
a large number of samples have to be tested. The tests give for brittle rocks, when tested under diametric loading,
values reasonable close to the BTS. Also it is suggested that PLS=0.8*BTS, it is suggested to establish such a
relation based on both tests.

Page 46 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Basic Soil Mechanics.

Figure 1-38: The vane shear test (English.geocpt.es).

Figure 1-39: Shear vane and Torvane for soil testing (www.humboldtmfg.com).

Copyright © Dr.ir. S.A. Miedema TOC Page 47 of 304


Introduction Dredging Engineering.

Figure 1-40: Point load test facility (inside.mines.edu).

Figure 1-41: Brazilian splitting tension test. Figure 1-42: BTS zoomed.

Page 48 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Basic Soil Mechanics.

Figure 1-43: A BTS test after failure.

1.7. Nomenclature.
Gs Specific gravity -
ρs Density of the soil kg/m3
ρw Density of water kg/m3
g Gravitational constant (9.81 m/s2) m/s2
Mt Mass of the soil, total mass kg
Ms Mass of the solids kg
Mw Mass of the water kg
Ma Mass of the air kg
Vt Volume of the soil, total volume m3
Vs Volume of the solids m3
Vw Volume of the water m3
Va Volume of the air m3
ρt Density of the soil kg/m3
γt Unit weight of the soil N/ m3
g Gravitational constant (9.81 m/s2) m/s2
Dr Relative density -
e Current void ratio of the soil in-situ -
emax Void ratio of the soil at its loosest condition -
emin Void ratio of the soil at its densest condition -
n Porosity of the soil in-situ -
nmax Porosity of the soil at its loosest condition -
nmin Porosity of the soil in its densest condition -
Vv Volume of the voids/pores m3
Vs Volume of the solids/grains/particles m3
n Porosity -
e Void ratio -
Ct Sorting coefficient -
C Sorting coefficient -
K Hydraulic conductivity m2

Copyright © Dr.ir. S.A. Miedema TOC Page 49 of 304


Introduction Dredging Engineering.

k Permeability m/s
f(n) porosity function -
C sorting coefficient
de effective grain diameter mm
d10 Grain diameter where 10% is smaller mm
d60 Grain diameter where 60% is smaller mm
U Grain uniformity coefficient -
v kinematic viscosity
μ Dynamic viscosity Pa.s
ρw Water density kg/m3
γw Unit weight of water N/m3
Q units of volume per time m³/s
K intrinsic permeability m2
k permeability m/s
A cross sectional area m2
L Length m
ua Start excess pore pressure Pa
ub End excess pore pressure Pa
μ dynamic viscosity of the fluid Pa.s
c Cohesion kPa
UCS Unconfined Compressive Strength kPa
V The total volume of soil m3
ni Initial porosity -
ncv Porosity at constant volume -
ε Dilatation -
c Unconfined Compressive Strength kPa
F Maximum Failure Load kN
A Cross-sectional area of the core sample m2
E Deformation modulus N/m2
W Specific work of failure J/m3
T Brazilian Tensile Strength (BTS) kPa
D Diameter of the core sample m
F Maximum Failure Load kN
L Length of the core sample m
IS Point load index kPa
F Failure load kN
De Distance between platen tips m
D e2 = D2 for diametrical test m2
D e2 = 4A/ = for axial, block and lump test m2
A = W.D = minimum cross-sectional area of a plane through the platen contact m2
points
ρw Density of water kg/m3
zw Depth below the water table m
u Hydrostatic pressure kPa
g Gravitational constant m/s2
σ1 the major principal stress kPa
σ3 the minor principal stress kPa
τ the shear strength τ = Su (or sometimes cu) kPa
Su the undrained strength kPa
σ' (σ – u) the effective stress kPa
σ Total stress applied normal to the shear plane kPa
u Pore water pressure acting on the same plane kPa
φ Effective stress friction angle or the angle of internal friction after Coulomb deg
friction
c' Cohesion kPa
τ The shear strength τ = Su (or sometimes cu) kPa

Page 50 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Basic Soil Mechanics.

1.8. Notes.

Copyright © Dr.ir. S.A. Miedema TOC Page 51 of 304


Introduction Dredging Engineering.

Page 52 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Basic Soil Mechanics.

Copyright © Dr.ir. S.A. Miedema TOC Page 53 of 304


Introduction Dredging Engineering.

Page 54 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Basic Soil Mechanics.

Copyright © Dr.ir. S.A. Miedema TOC Page 55 of 304


Introduction Dredging Engineering.

Page 56 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Terminal Settling Velocity of Particles.

Chapter 2: The Terminal Settling Velocity of Particles.

2.1. Introduction.
In many cases in hydraulic transport there will be equilibrium between erosion and deposition. In order to
understand this, both deposition and erosion (initiation of motion) will be discussed. The settling velocity of grains
depends on the grain size, shape and specific density. It also depends on the density and the viscosity of the carrier
liquid the grains are settling in, it also depends upon whether the settling process is laminar or turbulent.
Most slurry transport models use the terminal settling velocity, the particle drag coefficient or the particle Froude
number. So it is important to have a good unserstanding of these parameters.

2.2. The Equilibrium of Forces.


The settling velocity of grains depends on the grain size, shape and specific density. It also depends on the density
and the viscosity of the liquid the grains are settling in, and it depends upon whether the settling process is laminar
or turbulent. Discrete particles do not change their size, shape or weight during the settling process (and thus do
not form aggregates). A discrete particle in a liquid will settle under the influence of gravity. The particle will
accelerate until the frictional drag force of the liquid equals the value of the gravitational force, after which the
vertical (settling) velocity of the particle will be constant (Figure 2-1), the so called terminal settling velocity.

Figure 2-1: Forces on a settling particle.

The upward directed force on the particle, caused by the frictional drag of the liquid, can be calculated by:

1 2
Fu p  C D  l  vt  A (2-1)
2

The downward directed force, caused by the difference in density between the particle and the water can be
calculated by:

Fd o w n  (  s   l )  g  V   (2-2)

In this equation a shape factor ψ is introduced to compensate for the shape of real sand grains. This shape factor
is 1 for spheres and about 0.7 for real sand particles. The projected surface of the particle is:

 2
A  d (2-3)
4

The volume of the particle is:

Copyright © Dr.ir. S.A. Miedema TOC Page 57 of 304


Introduction Dredging Engineering.

 3
V  d (2-4)
6

In general, the terminal settling velocity vt can be determined with the following equation:

4  g  s   l  d  
vt  (2-5)
3lC D

The Reynolds number of the settling process determines whether the process is laminar or turbulent. The Reynolds
number can be determined by:

vt d
R ep  (2-6)
l

2.3. The Drag Coefficient.


In equation (2-5) all parameters are assumed to be known, except for the drag coefficient CD.

Drag coefficient of spheres


100

10
CD

0.1
1 10 100 1000 10000 100000
Re
Turton & Levenspiel Turton & Levenspiel Stokes

Figure 2-2: Experimental data for drag coefficients of spheres as a function of the Reynolds number
(Turton & Levenspiel, 1986).

The drag coefficient CD for spheres depends upon the Reynolds number according to:

The laminar region:

24
R ep  1  CD  (2-7)
R ep

Page 58 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Terminal Settling Velocity of Particles.

The transitional region:

24 3
1  R ep  2000  CD    0 .3 4 (2-8)
R ep R ep

The turbulent region:

R ep  2000  C D  0 .4 4 5 (2-9)

As can be seen from the above equations, the drag coefficient CD is not continuous at the transition points of Rep=1
and Rep=2000. To get a smooth continuous curve the following equations can be applied:

The laminar region:


24 3 24
R ep  1  C D  R ep  (   0 .3 4 )  (1 -R e p )  (2-10)
R ep R ep R ep

The transitional region:


24 3
1  R ep  2000  CD    0 .3 4 (2-11)
R ep R ep

The turbulent region:


10000 24 3 10000
R ep  10000  CD  (   0 .3 4 )  (1  )  0 .4 4 5 (2-12)
R ep R ep R ep R ep

R e ynolds num b er as a functio n of the p article d iam e te r


100000

10000

1000
R e y n o ld s n u m b e r

100

10

0 .1

0 .0 1
0 .0 1 0 .1 1 10 100

P a rtic le d ia m e te r (m m )

Figure 2-3: The particle Reynolds number as a function of the particle diameter.

Figure 2-3 shows the particle Reynolds number as a function of the particle diameter for sands and gravels, using
the Ruby & Zanke (1977) equation.

Another equation for the transitional region has been derived by Turton & Levenspiel (1986):

Copyright © Dr.ir. S.A. Miedema TOC Page 59 of 304


Introduction Dredging Engineering.

24 0 .6 5 7 0 .4 1 3
C   (1  0 .1 7 3  R e p ) (2-13)
D  1 .0 9
R ep 1  16300  R ep

The models derived to describe the Shields curve use the drag coefficient of spheres and hardly any discussion
about this has been found in literature, although it is known that for sands and gravels the drag coefficients,
especially at large Reynolds numbers, are larger than the drag coefficient for spheres. Engelund & Hansen (1967)
found the following equation based on measurements and found it best suited for natural sands and gravels (Julien,
1995):

24
C D   1 .5 (2-14)
R ep

It must be noted here that in general the drag coefficients are determined based on the terminal settling velocity of
the particles. Wu & Wang (2006) recently gave an overview of drag coefficients and terminal settling velocities
for different particle Corey shape factors. The result of their research is reflected in Figure 2-4. Figure 2-4 shows
the drag coefficients as a function of the Reynolds number and as a function of the Corey shape factor. Figure 2-5
shows the drag coefficient for natural sands and gravels. The asymptotic value for large Reynolds numbers is about
1, while equation (2-14) shows an asymptotic value of 1.5.
For shells lying flat on the bed, the drag coefficient will be similar to the drag coefficient of a streamlined half
body (0.09), which is much much smaller than the drag coefficient for settling (3). So there is a large asymmetry
between the settling process and the erosion process of shells, while for more or less spherical sand particles the
drag coefficient is considered to be the same in each direction.
Figure 2-6 shows the CD coefficient as a function of the Rep number. In the transition area the equations are
implicit. Iteration 1 shows the resulting CD values based on equations (2-7), (2-8) and (2-9), while iteration 2 shows
the results based on equations (2-10), (2-11) and (2-12). It is clear from this figure that iteration 2 matches the
observed data better than iteration 1, but equation (2-13) of Turton & Levenspiel (1986) matches the best. This is
however for spheres and not for real sand and gravel particles.

The drag coefficient for different shape factors


2
10

1
10
CD

0
10

-1
10
0 1 2 3 4 5 6
10 10 10 10 10 10 10
Re
Sf=1.0 Sf=0.80-0.99 Sf=0.6-0.79 Sf=0.4-0.59 Sf=0.20-0.39

Sf=1.0 Sf=0.9 Sf=0.7 Sf=0.5 Sf=0.3

Stokes

Figure 2-4: Drag coefficient as a function of the particle shape (Wu & Wang, 2006).

Page 60 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Terminal Settling Velocity of Particles.

The drag coefficient of natural sands


4
10

3
10

2
10
CD

1
10

0
10

-1
10
-3 -2 -1 0 1 2 3
10 10 10 10 10 10 10
Re
Wu & Wang Wu & Wang Stokes Julien

Figure 2-5: Drag coefficient for natural sediments (S f=0.7) (Wu & Wang, 2006).

2.4. Terminal Settling Velocity Equations.


Stokes, Budryck and Rittinger used these drag coefficients to calculate settling velocities for laminar settling
(Stokes), a transition zone (Budryck) and turbulent settling (Rittinger) of real sand grains. This gives the following
equations for the settling velocity:

Laminar flow, d<0.1 mm, according to Stokes.

2
v t  424  R sd d (2-15)

Transition zone, d>0.1 mm and d<1 mm, according to Budryck.

 (1  9 5  R sd  d 3)  1  (2-16)
v t  8 .9 2 5 
d

Turbulent flow, d>1 mm, according to Rittinger.

v t  8 7  R sd  d (2-17)

With the relative submerged density Rsd defined as:

s  l
R sd  (2-18)
l

In these equations the grain diameter is in mm and the settling velocity in mm/sec. Since the equations were derived
for sand grains, the shape factor for sand grains is included for determining the constants in these equations.

Another equation for the transitional region (in m and m/sec) has been derived by Ruby & Zanke (1977):

Copyright © Dr.ir. S.A. Miedema TOC Page 61 of 304


Introduction Dredging Engineering.

 3 
10   l R sd  g  d
vt   1  1 (2-19)
d  100   l
2 
 

The effective drag coefficient can now be determined by:

4 g R sd d 
C D   (2-20)
3 2
vt

Figure 2-7 shows the settling velocity as a function of the particle diameter for the Stokes, Budryck, Rittinger &
Zanke equations.

In these equations the grain diameter is in mm and the settling velocity in mm/sec, except for the Zanke equation.
Since the equations were derived for sand grains, the shape factor for sand grains is used for determining the
constants in the equations. The shape factor can be introduced into the equations for the drag coefficient by dividing
the drag coefficient by a shape factor . For normal sands this shape factor has a value of 0.7. The viscosity of the
water is temperature dependent. If a temperature of 10 is used as a reference, then the viscosity increases by 27%
at 0 and it decreases by 30% at 20 centigrade. Since the viscosity influences the Reynolds number, the settling
velocity for laminar settling is also influenced by the viscosity. For turbulent settling the drag coefficient does not
depend on the Reynolds number, so this settling process is not influenced by the viscosity. Other researchers use
slightly different constants in these equations but, these equations suffice to explain the basics of the Durand
theory.

C d a s a fu n ctio n o f th e R e n u m b e r
1000

100
Cd

10

0.1
0.1 1 10 100 1000 10000 100000

R e num ber
Iteration 1 Huisman Observed Iteration 2 Turton

Figure 2-6: The drag coefficient as a function of the particle Reynolds number.

Page 62 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Terminal Settling Velocity of Particles.

S ettlin g velo city o f rea l san d particles


1000

100
S e ttl i n g v e l o c i ty i n m m /s e c

10

0 .1
0 .0 1 0 .1 1 10

G ra in s iz e in m m
Sto ke s Bu d ryck Rittin g e r Z a n ke

Figure 2-7: The settling velocity of individual particles.

The Huisman (1973-1995) Method.

A better approximation and more workable equations for the drag coefficient CD may be obtained by subdividing
the transition region, for instance:

24
R ep  1 CD  (2-21)
1
R ep

24
1  R ep  50 CD  (2-22)
3/4
R ep

4 .7
50  R ep  1620 CD  (2-23)
1/3
R ep

1620  R ep C D  0 .4 (2-24)

This power approximation is also shown in Figure 2-6. Substitution of these equations in equation (2-19)
gives:

1
1 g 1 2
R ep  1 vt    R sd  d (2-25)
1
18 l

0 .8
1 g 0 .8 1 .4
1  R ep  50 vt    R sd  d (2-26)
0 .6
10 l

Copyright © Dr.ir. S.A. Miedema TOC Page 63 of 304


Introduction Dredging Engineering.

0 .6
1 g 0 .6 0 .8
50  R ep  1620 vt    R sd  d (2-27)
0 .2
2 .1 3 l

0 .5
g 0 .5 0 .5
1620  R ep v t  1 .8 3   R sd  d (2-28)
0
l

These equations are difficult to use in an actual case because the value of Rep depends on the terminal
settling velocity. The following method gives a more workable solution.

Equation (2-5) can be transformed into:

2 4 g 3
C D  R ep  R sd  d (2-29)
2
3 l

This factor can be determined from the equations above:

R ep  1 2
C D  R ep  24  R ep (2-30)

1  R ep  50 2 5/4
C D  R ep  24  R ep (2-31)

50  R ep  1620 2 5/3
C D  R e p  4 .7  R e p (2-32)

1620  R ep 2 2
C D  R e p  0 .4  R e p (2-33)

From these equations the equation to be applied can be picked and the value of Rep calculated. The
settling velocity now follows from:

l
v t  R ep  (2-34)
d

The Grace Method (1986).

Following the suggestions of Grace (1986), it is found convenient to define a dimensionless particle
diameter, which in fact is the Bonneville parameter (d in m and vt in m/s):

1/3
 R g 
sd
D*  d   (2-35)
  2 
 l 

And a dimensionless terminal settling velocity:

1/3
*
 1 
vt  vt   (2-36)

 l  R sd  g 

Those are mutually related. Thus using the curve and rearranging gives directly the velocity vt as a function of
particle diameter d. No iteration is required. This described by analytic expressions appropriate for a computational
determination of vs according to Grace Method. Now vt can be computed according to:

1/3
*
 1 
vt  vt   (2-37)
  l  R sd  g 

Page 64 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Terminal Settling Velocity of Particles.

D 
2
*

   
5 8
* 4 * 6 *
vt   3 .1 2 3 4  1 0  D  1 .6 4 1 5  1 0  D
D
*
 3 .8 18 (2-38)

 
11
10 *
 7 .2 7 8  1 0  D

* * 2
* *  1 .5 4 4 6  2 .9 1 6 2  lo g (D )  1 .0 4 3 2  lo g (D ) (2-39)
3 .8  D  7 .5 8 v t  10

* * 2 * 3
* *  1 .6 4 7 5 8  2 .9 4 7 8 6  lo g (D )  1 .0 9 7 0 3  lo g (D )  0 .1 7 1 2 9  lo g (D ) (2-40)
7 .5 8  D  227 v t  10

* * 2 * 3
* * 5 .1 8 3 7  4 .5 1 0 3 4  lo g (D )  1 .6 8 7  lo g (D )  0 .1 8 9 1 3 5 l o g (D ) (2-41)
227  D  3500 v t  10

Figure 2-8 shows the terminal settling velocity for the iterative method according to equations (2-10), (2-11) and
(2-12) and the methods of Huisman (1973-1995) and Grace (1986), using shape factors of 0.5 and 0.7. It can be
seen that for small diameters these methods gives smaller velocities while for larger diameters larger velocities are
predicted, compared with the other equations as shown in Figure 2-7. The iterative method gives larger velocities
for the larger diameters, compared with the Huisman and Grace methods, but this is caused by the different way
of implementing the shape factor. In the iterative method the shape factor is implemented according to equation 2,
while with the Huisman and Grace methods the terminal settling velocity for spheres is multiplied by the shape
factor according to equation (2-42). For the smaller grain diameters, smaller than 0.5 mm, which are of interest
here, the 3 methods give the same results.

S ettlin g velo city o f pa rticle s u sing the sh ape factor


1000

100
S e ttlin g v e lo c ity in m m /s e c

10

0 .1
0 .0 1 0 .1 1 10

G ra in s iz e in m m
Ite ra tio n (0 .7 ) H u isma n (0 .7 ) G ra ce (0 .7 ) Ite ra tio n (0 .5 ) H u isma n (0 .5 ) G ra ce (0 .5 )

Figure 2-8: The settling velocity of individual particles using the shape factor.

2.5. The Shape Factor


In the range of particle Reynolds numbers from roughly unity to about 100, which is the range of interest here, a
particle orients itself during settling so as to maximize drag. Generally this means that an oblate or lenticular
particle, i.e. a shape with one dimension smaller than the other two, will settle with its maximum area horizontal.
The drag of fluid on the particle then depends most critically on this area. This is also the area seen if the particle
lies in a stable position on a flat surface. Therefore, for estimation of drag, the non-spherical particle is

Copyright © Dr.ir. S.A. Miedema TOC Page 65 of 304


Introduction Dredging Engineering.

characterized by the ‘area equivalent diameter’, i.e. the diameter of the sphere with the same projected area. For
particles whose sizes are determined by sieving rather than microscopic analysis, the diameter is slightly smaller
than the mesh size. However, unless the particles are needle shaped, the difference between the diameter and the
screen opening is relatively small, generally less than 20%.

Although equation (2-5) contains a shape factor, basically all the equations in this chapter are derived for spheres.
The shape factor ψ in equation (2-5) is one way of introducing the effect of the shape of particles on the terminal
settling velocity. In fact equation (2-5) uses a shape factor based on the weight ratio between a real sand particle
and a sphere with the same diameter. Another way is introducing a factor ξ according to:

vt
  (2-42)
v ts

Where ξ equals the ratio of the terminal settling velocity of a non-spherical particle vt and the terminal velocity vts
of a spherical particle with the same diameter. The shape of the particle can be described by the volumetric shape
factor K which is defined as the ratio of the volume of a particle and a cube with sides equal to the particle diameter
so that K=0.524 for a sphere:

v o lu m e of p a r tic le (2-43)
K 
3
d

The shape factor ξ is a function of the volumetric form factor K and the dimensionless particle diameter D*
according to equation (2-35).

 0 .6 K  0 .5 2 4
K  0 .5 2 4  0 .0 4 5  0 .0 5  K  0 .0 2 8 7  5 5 0 0 0
lo g (  )   0 .5 5  K  0 .0 0 1 5  K  0 .0 3  1 0 0 0  (2-44)
*
c o s h ( 2 .5 5  (lo g (D )  1 .1 1 4 )

This equation takes a simpler form for sand shaped particles with K=0.26:

0 .0 6 5 6
lo g (  )   0 .3 0 7 3  (2-45)
*
c o s h ( 2 .5 5  ( lo g ( D )  1 .1 1 4 )

A value of K=0.26 for sand grains would give a volume ratio of 0.26/0.524=0.496 and thus a factor ψ=0.496 in
equation (2-5), while often a factor ψ=0.7 is used.

Figure 2-9 shows the shape factor ξ as a function of the dimensionless particle diameter d*, according to equation
(2-44).

Figure 2-8 also shows the terminal settling velocity according to the methods of Huisman (1973-1995) and Grace
(1986) using the shape factor according to equation (2-45). It can be seen clearly that both methods give the same
results. One can see that the choice of the shape factor strongly determines the outcome of the terminal settling
velocity.

Page 66 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Terminal Settling Velocity of Particles.

S h a p e fa c to r
1.0

0.9

0.8
S h a p e fa c to r

0.7

0.6

0.5

0.4

0.3

0.2
1 10 100

D im e n s io n le s s p a rtic le d ia m e te r d *
K=0.26 K=0.1 K=0.2 K=0.3 K=0.4 K=0.5

Figure 2-9: The shape factor ξ as a function of the dimensionless particle diameter D*.

2.6. Hindered Settling.


The above equations calculate the settling velocities for individual grains. The grain moves downwards and the
same volume of water has to move upwards. In a mixture, this means that, when many grains are settling, an
average upwards velocity of the water exists. This results in a decrease of the settling velocity, which is often
referred to as hindered settling. However, at very low concentrations the settling velocity will increase because the
grains settle in each other’s shadow. Richardson and Zaki (1954) determined an equation to calculate the influence
of hindered settling for volume concentrations Cvs between 0 and 0.3. The coefficient in this equation is dependent
on the Reynolds number. The general equation yields:

v th 
 1  C vs  (2-46)
vt

The following values for  should be used according to Richardson and Zaki (1954):

Rep<0.2 =4.65
Rep>0.2 and Rep<1.0 =4.35Rep-0.03
(2-47)
Rep>1.0 and Rep<200 =4.45Rep-0.1
Rep>200 =2.39

However this does not give a smooth continuous curve. Using the following definition does give a continuous
curve:

Rep<0.1 =4.65
Rep>0.1 and Rep<1.0 =4.35Rep-0.03
(2-48)
Rep>1.0 and Rep<400 =4.45Rep-0.1
Rep>400 =2.39

Copyright © Dr.ir. S.A. Miedema TOC Page 67 of 304


Introduction Dredging Engineering.

Other researchers found the same trend but sometimes somewhat different values for the power β. These equations
are summarized below and shown in Figure 2-10.

According to Rowe (1987) this can be approximated by:

0 .7 5
4 .7  0 .4 1  R e p
  (2-49)
0 .7 5
1  0 .1 7 5  R e p

Wallis (1969) found an equation which matches Rowe (1987) for small Reynolds numbers and Garside & Al-
Dibouni (1977) for the large Reynolds numbers:

0 .6 8 7
4 .7  (1  0 .1 5  R e p )
  (2-50)
0 .6 8 7
1  0 .2 5 3  R ep

Garside & Al-Dibouni (1977) give the same trend but somewhat higher values for the exponent β.

0 .9
5 .1  0 .2 7  R e p
  (2-51)
0 .9
1  0 .1  R e p

Di Felici (1999) finds very high values for β but this relation is only valid for dilute mixtures (very low
concentration, less than 5%).

0 .7 4
6 .5  0 .3  R e p
  (2-52)
0 .7 4
1  0 .1  R e p

Hindered settling power


7

4
Beta (-)

0
10-3 10-2 10-1 100 101 102 103 104
Re p (-)
Rowe Wallis Garside Di Felici Richardson & Zaki

Figure 2-10: The hindered settling power according to several researchers.

Page 68 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Terminal Settling Velocity of Particles.

2.7. Conclusions.
The equation of Ruby & Zanke (1977) will be used to determine the terminal settling velocity for sands and gravels.
The equation of Richardson and Zaki (1954) will be used for hindered settling, with the equation of Rowe (1987)
for the power β in the hindered settling equation. The DHLLDV framework is calibrated based on these equations.

Using different equations will result in slightly different hydraulic gradients and Limit Deposit Velocities,
requiring the constants in the DHLLDV framework to be adjusted.

Particles with different shapes, like spheres or shells, and particles with different relative submerged densities may
require different methods.

Copyright © Dr.ir. S.A. Miedema TOC Page 69 of 304


Introduction Dredging Engineering.

2.8. Nomenclature.
A Cross section of particle m2
CD Drag coefficient -
Cvs Volumetric spatial concentration -
d Particle diameter m
D* Bonneville parameter or dimensionless particle diameter -
Fdown Downwards force on particle N
Fup Upwards force on particle N
g Gravitational constant 9.81 m/s2 m/s2
K Volumetric form factor -
Rep Particle Reynolds number -
Rsd Relative submerged density -
vt Terminal settling velocity m/s
v t* Dimensionless terminal settling velocity -
vth Hindered terminal settling velocity m/s
vts Terminal settling velocity sphere m/s
V Volume of particle m3
β Hindered settling power -
ρl Density of carrier liquid ton/m3
ρs Density of solid ton/m3
ψ Shape factor -
ξ Shape factor -
νl Kinematic viscosity liquid m2/s

Page 70 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Terminal Settling Velocity of Particles.

2.9. Notes.

Copyright © Dr.ir. S.A. Miedema TOC Page 71 of 304


Introduction Dredging Engineering.

Page 72 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Terminal Settling Velocity of Particles.

Copyright © Dr.ir. S.A. Miedema TOC Page 73 of 304


Introduction Dredging Engineering.

Page 74 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Terminal Settling Velocity of Particles.

Copyright © Dr.ir. S.A. Miedema TOC Page 75 of 304


Introduction Dredging Engineering.

Page 76 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Pump/Pipeline System.

Chapter 3: The Pump/Pipeline System.

3.1. Introduction.
A multi pump/pipeline system consists of components with different dynamic behaviour. To model such a system,
one should start with simple mathematical descriptions of the sub-systems, to be able to determine the sensitivity
of the behaviour of the system to changes in one of the sub-systems. The following sub-systems can be
distinguished:
- The pump drive
- The centrifugal pump
- The sand/water slurry in the pipeline
- Flow control (if used)

Figure 3-1: A pump –pipeline system.

The system is limited by cavitation at the entrance of each pump on one hand and by sedimentation of the solids
resulting in plugging of the pipeline on the other hand. Cavitation will occur at high line velocities and/or at high
solids concentrations in the suction pipe of the pump considered. Sedimentation will occur at line velocities below
the so called critical velocity. The critical velocity depends on the grain distribution and on the solids concentration.
In between these two limitations a stable transportation process is required. A steady state process is possible only
if the solids properties and the solids concentration are constant in time. In practice however this will never be the
case. Solids properties such as the grain size distribution will change as a function of time and place as will the
solids concentration. The resistance of the slurry flow depends on the solids properties and concentration. If the
total resistance of the slurry flow in a long pipeline is considered, changes of the solids properties and concentration
at the suction mouth will result in slow changes of the total resistance, since only a small part of the pipeline is
filled with the new slurry while most of the pipeline remains filled with the slurry that was already there, except
from the slurry that has left the pipeline at the end. If the relatively short suction line is considered, this results in
a much faster change of the vacuum at the inlet of the first pump.
The total head of a pump however, responds immediately to changes of the solids properties and concentration. If
a sudden increase of the concentration is assumed, the total head of a pump will increase almost proportionally
with the concentration. This will result in a higher flow velocity, but, because of the inertia of the slurry mass in
the pipeline, the slurry mass will have to accelerate, so the flow velocity responds slowly on changes of the total
head. The increase of the total head also causes an increase of the torque and power of the pump drive, resulting
in a decrease of the pump drive revolutions and thus of the total head. Because of the inertia of pump and pump
drive, there will not be an immediate response.
It is obvious that there is an interaction between all the different sub-systems. These interactions can be ranged
from very slow to immediate. To be able to model the system, first the characteristic behaviour of the sub-systems
should be known.

Copyright © Dr.ir. S.A. Miedema TOC Page 77 of 304


Introduction Dredging Engineering.

3.2. The Pump Drive.


Pump drives used in dredging are diesel direct drives, diesel/electric drives and diesel/hydraulic drives. In this
paper the diesel direct drive, as the most common arrangement, is considered.
At nominal operating speed, the maximum load coincides with the nominal full torque point. If the torque is less
then the nominal full torque, the engine speed usually rises slightly as the torque decreases. This is the result of
the control of the speed by the governor. The extent of this depends upon the type of governor fitted.
If the engine load increases above the full torque point, the speed decreases and the engine operates in the full fuel
range. With most diesel engines the torque will increase slightly as the speed decreases, because of a slightly
increasing efficiency of the fuel pumps. When the load increases further, insufficient air is available to produce
complete combustion and the engine stalls. The torque drops rapidly and heavily polluted gasses are emitted. The
smoke limit has been reached. The speed range between the full torque point and the smoke limit is often referred
to as the constant torque range.

Figure 3-2: The speed-power curve of a diesel engine.

The torque/speed characteristic of the diesel engine can thus be approximated by a constant full torque upon the
nominal operating speed, followed by a quick decrease of the torque in the governor range.
This characteristic however is valid for a steady state process of the diesel engine. When the speed of the diesel
changes, the load will change, but also the inertia effects of the diesel have to be taken into account. The equation
of motion of the diesel engine, gear box and centrifugal pump combination, reduced to the axis of the centrifugal
pump, is:

 I d .e .  I g .b .  I c .p .     T d .e .  T h .t . (3-1)

In a steady state situation, the torque delivered by the diesel engine Td.e. equals the torque required by the
hydraulic transport Th.t., so the angular acceleration of the diesel is zero. If Td.e. is greater then Th.t., the
revolutions will increase, If Td.e. is smaller then Th.t., the revolutions will decrease. If the difference between
these two torque's is approximated to be proportional with the difference between the actual angular velocity and
the nominal operating angular velocity:

T d .e .  T h .t .  K p 
  s .p .    (3-2)

Page 78 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Pump/Pipeline System.

The linear differential equation can be written as:

 I d .e .  I g .b .  I c .p .     K p 
  s .p .    (3-3)

K
 I d .e .  I g .b .  I c .p .  
p
With: It and  d .e .  (3-4)
It

The solution of this first order system is:

  0    s .p .   0    1  e  t /   d .e .
(3-5)

In which  0 is the angular velocity at an arbitrary time, defined as t=0. Using time domain calculations with a
time step  t , the angular velocity at time step n can now be written as a function of the angular velocity at time
step n-1 and the set point angular velocity  s .p . according to:

 n   n1    s .p .   n  1    1  e   t /   d .e .
(3-6)

3.3. The Centrifugal Pump.

Figure 3-3: Centrifugal pumps.

Copyright © Dr.ir. S.A. Miedema TOC Page 79 of 304


Introduction Dredging Engineering.

The behaviour of centrifugal pumps can be described with the Euler impulse moment equation:

 Q  c o t(  o )   Q  c o t(  i ) 
pE  f  uo   uo    f  ui   ui   (3-7)
 2    ro   2    ri 

For a known pump this can be simplified to:

pE  f  C1  C 2  Q  (3-8)

Figure 3-4: The pressure-flow curves.

Because of incongruity of impeller blades and flow, the finite number of blades, the blade thickness and the internal
friction of the fluid, the Euler pressure  p E has to be corrected with a factor k, with a value of about 0.8. This
factor however does not influence the efficiency. The resulting equation has to be corrected for losses from
frictional contact with the walls and deflection and diversion in the pump and a correction for inlet and impact
losses. The pressure reduction for the frictional losses is:

2
 p h .f .  C 3   f  Q (3-9)

For a given design flow Qd the impact losses can be described with:

2
 p h .i .  C 4   f   Q d  Q  (3-10)

The total head of the pump as a function of the flow is now:

 
2
 p p  k   p E   p h .f .   p h . i .   f  k   C 1  C 2  Q   C4  Q d  Q 
2
C3 Q (3-11)

This is a second degree polynomial in Q. The fluid density  f in the pump can be either the density of a
homogeneous fluid (for water  w ) or the density of a mixture  m passing the pump.
The total efficiency of the pump can be determined by dividing the power that is added to the flow P f l   p p  Q
by the power that is output of the diesel engine Pd .e .  k   p E  Q  Pd .f . (in which P d .f . is the power required for
the frictional losses in the gear box, the pump bearings, etc.), this gives:

Page 80 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Pump/Pipeline System.

( k   p E   p h .f .   p h . i . )  Q
p  (3-12)
k   p E  Q  P d .f .

For the efficiency curve a third degree polynomial approximation satisfies, while the power and torque curves
approximate straight lines. The pump characteristics usually will be measured for a specific impeller diameter and
number of revolutions.

3.4. Affinity Laws.


In a dynamic system however, the pump revolutions will change. This is on one hand the result of the torque/speed
curve of the pump drive, on the other hand of manual or automatic flow control. This means that the pump
characteristics should also be known at different pump speeds.
The so-called affinity laws describe the influence of a different impeller diameter or revolutions on the pump head,
flow and efficiency:

2 2 2
p1 n1 D1 Q1 n1 D1
  ,   , 1   2 (3-13)
2 2 2
p2 n2 D2 Q2 n2 D2

The efficiency does not change, but the value of the flow on the horizontal axis is shifted.
The affinity laws for the power and the torque can easily be derived from these equations.

3 4 2 4
P1 p1  Q 1  2 n1 D1 T1 P1  n 2 p1  Q 1  2  n 2 n1 D1
   ,     (3-14)
P2 p 2  Q 2  1 3
n2
4
D2 T2 P2  n 1 p 2  Q 2  1  n 1 2
n2
4
D2

n D
If a ratio for the revolutions n  and a ratio for the diameter D  are given, the head and efficiency
nm Dm
curves at a speed n and an impeller diameter D can be determined by:

1 2
Q  Q m  n  D (3-15)

0 2 2 1 1 0 2 0 2
pp  0  Q  n  D  1  Q  n  D   2  Q  n  D (3-16)

0 0 0 1 1 2 2 2 4 3 3 6
p   0  Q  n  D  1  Q  n  D   2  Q  n  D   3  Q  n  D (3-17)

In which nm, Dm and Qm are the revolutions, impeller diameter and flow used in the measurements of the head
and efficiency curves.
Based on this theory, the characteristics of two pumps used in the case study in this chapter, are given in Figure
3-5 and Figure 3-6. Both pumps are limited by the constant torque behaviour of the corresponding diesel engine
in the full fuel range.

Figure 3-5 and Figure 3-6 give the pump characteristics for clear water. If a mixture is pumped however, the pump
head increases because of the mixture density as has been pointed out when discussing equation (3-11) and the
pump efficiency decreases because a heterogeneous mixture is flowing through the pump. The decrease of the
efficiency depends upon the average grain diameter, the impeller diameter and the solids concentration and can be
determined with (according to Stepanoff):

m   1  C t   0 .4 6 6  0 .4  L o g 1 0  d 5 0   / D  (3-18)

Copyright © Dr.ir. S.A. Miedema TOC Page 81 of 304


Introduction Dredging Engineering.

Pum p H ead Pum p Pow er


500 1700

400 1360

300 1020
k P a

k W
200 680

100 340

0 0
0 .0 0 0 .4 0 0 .8 0 1 .2 0 1 .6 0 2 .0 0 2 .4 0 2 .8 0 3 .2 0 3 .6 0 4 .0 0 0 .0 0 0 .4 0 0 .8 0 1 .2 0 1 .6 0 2 .0 0 2 .4 0 2 .8 0 3 .2 0 3 .6 0 4 .0 0
F lo w in m ^ 3 / s e c F lo w in m ^ 3 / s e c

P u m p E ffic ie n c y P u m p T o rq u e
1 .0 0 0 60

0 .8 0 0 48

0 .6 0 0 36

k N m
-

0 .4 0 0 24

0 .2 0 0 12

0 .0 0 0 0
0 .0 0 0 .4 0 0 .8 0 1 .2 0 1 .6 0 2 .0 0 2 .4 0 2 .8 0 3 .2 0 3 .6 0 4 .0 0 0 .0 0 0 .4 0 0 .8 0 1 .2 0 1 .6 0 2 .0 0 2 .4 0 2 .8 0 3 .2 0 3 .6 0 4 .0 0
F lo w in m ^ 3 / s e c F lo w in m ^ 3 / s e c

Pum p N PSH P u m p R e v o lu t io n s
100 400

80 320

60 240
k P a

rp m
40 160

20 80

0 0
0 .0 0 0 .4 0 0 .8 0 1 .2 0 1 .6 0 2 .0 0 2 .4 0 2 .8 0 3 .2 0 3 .6 0 4 .0 0 0 .0 0 0 .4 0 0 .8 0 1 .2 0 1 .6 0 2 .0 0 2 .4 0 2 .8 0 3 .2 0 3 .6 0 4 .0 0
F lo w in m ^ 3 / s e c F lo w in m ^ 3 / s e c

P u m p P ip e lin e S y s te m W in d o w s V 2 . 0 1
P u m p R e vo lu tio n s
0 5 - 1 2 - 1 9 9 6 - 1 0 :2 6 :2 5
In p u t: 2 5 5 rp m 2 0 5 rp m 2 3 0 rp m 2 5 5 rp m 2 8 0 rp m 3 0 5 rp m
P u m p F ile : C :\ P P S W \P u m p \L a d d e r . P m p
T o r q u e L im ite d

Figure 3-5: The characteristics of the ladder pump.

Pum p H ead Pum p Pow er


2100 7200

1680 5760

1260 4320
k P a

k W

840 2880

420 1440

0 0
0 .0 0 0 .4 0 0 .8 0 1 .2 0 1 .6 0 2 .0 0 2 .4 0 2 .8 0 3 .2 0 3 .6 0 4 .0 0 0 .0 0 0 .4 0 0 .8 0 1 .2 0 1 .6 0 2 .0 0 2 .4 0 2 .8 0 3 .2 0 3 .6 0 4 .0 0
F lo w in m ^ 3 / s e c F lo w in m ^ 3 / s e c

P u m p E ffic ie n c y P u m p T o rq u e
1 .0 0 0 180

0 .8 0 0 144

0 .6 0 0 108
k N m
-

0 .4 0 0 72

0 .2 0 0 36

0 .0 0 0 0
0 .0 0 0 .4 0 0 .8 0 1 .2 0 1 .6 0 2 .0 0 2 .4 0 2 .8 0 3 .2 0 3 .6 0 4 .0 0 0 .0 0 0 .4 0 0 .8 0 1 .2 0 1 .6 0 2 .0 0 2 .4 0 2 .8 0 3 .2 0 3 .6 0 4 .0 0
F lo w in m ^ 3 / s e c F lo w in m ^ 3 / s e c

Pum p N PSH P u m p R e v o lu t io n s
100 500

80 400

60 300
k P a

rp m

40 200

20 100

0 0
0 .0 0 0 .4 0 0 .8 0 1 .2 0 1 .6 0 2 .0 0 2 .4 0 2 .8 0 3 .2 0 3 .6 0 4 .0 0 0 .0 0 0 .4 0 0 .8 0 1 .2 0 1 .6 0 2 .0 0 2 .4 0 2 .8 0 3 .2 0 3 .6 0 4 .0 0
F lo w in m ^ 3 / s e c F lo w in m ^ 3 / s e c

P u m p P ip e lin e S y s te m W in d o w s V 2 . 0 1
P u m p R e vo lu tio n s
0 5 - 1 2 - 1 9 9 6 - 1 0 :2 7 :1 6
In p u t: 3 3 0 rp m 2 6 0 rp m 2 9 5 rp m 3 3 0 rp m 3 6 5 rp m 4 0 0 rp m
P u m p F ile : C :\ P P S W \P u m p \M a in . P m p
T o r q u e L im ite d

Figure 3-6: The characteristics of the main pump and the booster pump, torque limited.

Page 82 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Pump/Pipeline System.

3.5. Approximations.
The pump Q-H curve for water can be approximated by:

2 (1  f )  H D
H  A  C Q W ith : A = f  H D and C= (3-19)
2
Q D

The pump efficiency curve can be approximated by:

g
  QD  Q  
2

 m   m ax 
 1    (3-20)
  QD  

A factor f of about 1.25 and a power g of about 0.7 give good results. The design flow QD (m3/s) and the design
head HD (kPa) follow from the requirements and can be used for one pump or for all pumps together.

3.6. The Total Head Losses.


The pressure at the inlet of the suction mouth of the cutter head or in the draghead is:

p s   w  g  H sm  1 0 0 (3-21)

The pressure losses from the suction mouth to the entrance of the first pump are:

n
Ls

1 2 1 2 1 2
p m ,s    m  v ls ,s   w ,s     w  v ls ,s  p s ,s   n ,s    m  v ls ,s   m  g  H s
2 2 2
D p ,s 1
(3-22)

+  m  L s  v ls ,s

The pressure losses after the first pump (discharge losses) are:

 
Ld

1 2 2 1 2 1 2
p m ,d    m  v ls ,d  v ls ,s w ,d     w  v ls ,d  p s ,d   n ,d    m  v ls ,d   m  g  H d
2 2 2
D p ,d 1
(3-23)

+  m  L d  v ls ,d

The absolute pressure at the inlet of the first pump is ps-pm,s and should be above a certain minimum.

p s  p m ,s  p lim (3-24)

The last term in equations (3-22) and (3-23) is the contribution of acceleration or deceleration of the mixture.

In a stationary situation, the mixture density has to be smaller than a certain limit to avoid cavitation in the first
pump.

w g  H sm  1 0 0  p lim
m 
n
Ls (3-25)

2 2 2
1
 v ls ,s   w ,s   1
 v ls ,s   n ,s  1
 v ls ,s  g  H s
2 2 2
D p ,s 1

Copyright © Dr.ir. S.A. Miedema TOC Page 83 of 304


Introduction Dredging Engineering.

Figure 3-7: Characteristics of the pump/pipeline system, not limited.

Figure 3-8: Characteristics of the pump/pipeline system, torque limited.

Page 84 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Pump/Pipeline System.

3.7. The Pump/Pipeline System Description.


In a steady state situation, the revolutions of the pumps are fixed, the line speed is constant and the solids properties
and concentration are constant in the pipeline. The working point of the system is the intersection point of the
pump head curve and the pipeline resistance curve. The pump curve is a summation of the head curves of each
pump according to equation (3-16). The resistance curve is a summation of the resistances of the pipe segments
and the geodetic head according to equations (3-22) and (3-23). Figure 3-7 and Figure 3-8 show this steady state
situation for the system used in the case study at 6 densities ranging from clear water upto a density of 1.6 ton/m3.
In reality, the solids properties and concentration are not constant in time at the suction mouth. As a result of this,
the solids properties and concentration are not constant as a function of the position in the pipeline. To be able to
know these properties as a function of the position in the pipeline, the pipeline must be divided into small segments.
These segments move through the pipeline with the line speed. Each time step a new segment is added at the
suction mouth, while part of the last segment leaves the pipeline. Because the line speed is not constant, the length
of the segment added is not constant, but equals the line speed times the time step. For each segment the resistance
is determined, so the resistance as a function of the position in the pipeline is known. This way also the vacuum
and the discharge pressure can be determined for each pump. If vacuum results in cavitation of one of the pumps,
the pump head is decreased by decreasing the pump density, depending on the time the pump is cavitating. The
dynamic calculations are carried out in the time domain, because most of the equations used are non-linear. The
time step used is about 1 second, depending on the speed of the PC and the other tasks Windows has to carry out.

3.8. The Segmented Pipeline.


In reality, the solids properties and concentration are not constant in time at the suction mouth. As a result of this,
the solids properties and concentration are not constant as a function of the position in the pipeline. To be able to
know these properties as a function of the position in the pipeline, the pipeline must be divided into small segments.
These segments move through the pipeline with the line speed. Each time step a new segment is added at the
suction mouth, while part of the last segment leaves the pipeline. Because the line speed is not constant, the length
of the segment added is not constant, but equals the line speed times the time step. For each segment the resistance
is determined, so the resistance as a function of the position in the pipeline is known. This way also the vacuum
and the discharge pressure can be determined for each pump. If vacuum results in cavitation of one of the pumps,
the pump head is decreased by decreasing the pump density, depending on the time the pump is cavitating.

As mentioned before, each segment contains the mixture properties. The two most important properties are the
mixture density and the grain size distribution. If a homogeneous transport model is considered, the grain
distribution can be replaced by the characteristic factor depending on the grain size distribution. For a
heterogeneous or two-phase transport model, the problem becomes much more complicated.

The segments move through the pipeline with the line speed, assuming that all of the contents of a segment move
at the same speed. However if part of the mixture has settled at the bottom of the pipeline, this part will move with
a much smaller velocity then the average velocity, while the mixture above the sediment will move with a velocity
higher then the average. In a stationary situation this does not matter, as long as the transport model used takes this
into account (the Durand model takes this into account), but in a non-stationary situation there may be temporary
accumulation of solids. Also dunes may occur, moving through the pipeline. To implement these phenomena a
longitudinal diffusion model has to be developed. The current administrative system in the simulation software is
suitable for storing the information required to describe these phenomena. However the information stored has to
be extended, since two-phase flow requires storage of two components, the bed load and the suspended material.
With a time step in the simulation software of 0.1 to 0.2 seconds, the segment length varies (with a line speed of 5
m/s) from 0.5 to 1.0 m. The required length for a good description of dunes moving through the pipeline is
unknown, but from experiments in our laboratory it seems a segment length of 0.5 m is still to high. An intuitive
estimate of 0.1 to 0.2 m seems reasonable. The Durand model however has not been developed for a pipeline of
only 0.1 m.

The mass conservation equation of a pipe segment can be described with equation (3-26). In this equation all terms
give a mass flow. The sum of the mass flow of the suspended material and the bed load that enter a segment, should
be equal to the sum of the suspended material and the bed load that leave the segment plus the material that settles
in the segment. The last term on the right hand side is the settlement of suspended material into the bed. This term
is positive when material settles (accumulates) in the segment.

Q in  s  Q in  b  Q out  s  Q out  b  Q s b (3-26)

Copyright © Dr.ir. S.A. Miedema TOC Page 85 of 304


Introduction Dredging Engineering.

Su sp e n d e d Q
Q in  s Q o ut  s
s b

Q in  b
Be d lo a d Q out  b

Pip e se g m e n t
Figure 3-9: The mass equilibrium in a pipe segment.

The question is however; whether for a good description of the transport it suffices to administer the suspended
load and the bed load in one segment moving through the pipeline. In fact the velocity of the suspended load will
be higher then the average line speed and the velocity of the bed will be much smaller. The pipe segment should
have to be split into two separate segments for the suspended load and for the bed load, moving at two different
velocities through the system, in order to administer the two phase flow correctly. The current method of
administering the contents of the segments is suitable for suspended load only at line speeds above the critical
velocity.
A good description of the vertical diffusion between the suspended load and the bed load is not yet available and
will be subject for further research. Erosion diffusion equations are used for hopper sedimentation as well, but
these equation do not suffice Miedema and Vlasblom (1996).

3.9. The Inertial Effects in the Pipeline.


A steady state process requires a constant density and solids properties in the system and thus at the suction mouth.
In practice it is known, that the solids properties and the density change with respect to time. As a result, the pump
discharge pressure and vacuum will change with respect to time and the pipeline resistance will change with respect
to time and place. A change of the discharge pressure will result in a change of the torque on the axis of the pump
drive on one hand and in a change of the flow velocity on the other hand. The mixture in the pipeline has to
accelerate or decelerate. Since centrifugal pumps respond to a change in density and solids properties at the
moment the mixture passes the pump, while the pipeline resistance is determined by the contents of the pipeline
as a whole, this forms a complex dynamic system.

The inertial pressure of the mixture has to be added to the resistance of the mixture. In fact, the inertial pressure is
always equal to the difference between the total pressure generated by the pumps and the total resistance of the
mixture in the pipeline system. If this difference is positive (the pump pressure has increased due to an increase of
the mixture density), the mixture will accelerate. If negative, the mixture will decelerate.
As a result of the acceleration and deceleration, the mixture velocity (line velocity) will vary as a function of time.
To realize a stable dredging process, it is required to have a line velocity that will not vary too much. The line
velocity can be controlled by varying the revolutions of one of the dredge pumps, where the last pump is preferred.
From the above one can distinguish the different effect by the time they require to change/occur:
1. Very fast (within a second), the change in discharge pressure of a centrifugal pump
2. Fast (seconds), the change in revolutions of the pump drive and the change in line speed (acceleration
and deceleration)
3. Slow (minutes), filling up the pipeline with mixture or a change in mixture content

These effects can also be recognized in the equations describing the pump curve and the system curve. Equation
(3-11) shows the effect of the fluid (mixture) density on the discharge pressure. Equation (3-6) shows the effect
of a changing set point of the pump drive. Equations (3-22) and (3-23) contain the inertial effect in the most right
term on the right hand side, while the effect of the changing mixture contents is described by the first term on the
right hand side. Figure 3-7 and Figure 3-8 show the system curves and the pump curves for the system described
in Figure 3-1, for 5 different densities, including clear water, for a stationary situation. The intersection points of
each system and pump curve at one density are the working points for the system at that specific density.

Page 86 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Pump/Pipeline System.

Figure 3-10: The system curves for 3 cases, accelerating.

Figure 3-10 is a representation of a number of phenomena that occur subsequently when the system (Figure 3-1)
filled with water, is filled with mixture with a density of 1.6 ton/m3. In this figure case 1 represents the system and
the pump curve for the system filled with water. Case 2 represents the system with the pipeline filled with mixture
up to a point just before the 3rd (booster) pump. Case 3 represents the system filled entirely with the mixture.

Now, what happens if a system filled with water is continuously filled with the mixture?

First the working point is point 1 in Figure 3-10. This is the intersection point of the pump and system curves for
water. When mixture enters the system, within a few (about 8) seconds the mixture has reached the ladder and
main pump, since the distance is only about 44 m and the line speed about 5 m/sec. At that moment, the discharge
pressure of the ladder pump and main pump increase proportionally to the mixture density, resulting in a pump
curve according to case 2 and a working point 2. The flow and thus the line speed will not change instantly because
of the inertia of the fluid and solids mass in the pipeline. Number 6 shows the access pressure caused by the sudden
increase of the discharge pressure. This access pressure has to take care of the acceleration of the pipeline contents.
This acceleration will take in the order of 10-20 seconds.

The filling of the system continues and the resistance of the mixture slowly increases, so the working point moves
from point 2 to point 3. With the line speed of 5 m/s, this will take about 400 seconds or almost 7 minutes. When
the mixture reaches the booster pump, at once the discharge pressure increases, resulting in the pump curve

Copyright © Dr.ir. S.A. Miedema TOC Page 87 of 304


Introduction Dredging Engineering.

according to case 3, the top curve. The working point will move to point 4, while 7 represents the access pressure
causing the acceleration of the pipeline contents. Moving from 3 to 4 will take 10-20 seconds. When the pipeline
continues to be filled with mixture, the resistance increases, resulting in the working point moving from 4 to 5 in
about 400 seconds.

Figure 3-11: The system curves for 3 cases, decelerating.

Figure 3-11 shows the same procedure for a pipeline filled with a mixture of density 1.6 ton/m 3. In this case the
pipeline, containing mixture of 1.6 ton/m3, is filled with water, resulting in decreasing discharge pressures and
pipeline resistance. The procedure is almost the inverse, but Figure 3-11 shows that the path followed is different.
In working point 1, all the pumps and the pipeline are filled with the mixture. When the water reaches the ladder
and main pump, the pump curve is decreased to case 2 and the new working point is point 2. 6 gives the deceleration
pressure, so the contents of the pipeline will decelerate from 1 to 2 in about 10-20 seconds. From 2-3 the pipeline
is filled with water up to the booster pump, resulting in a decrease of the resistance, taking about 400 seconds.
When the water reaches the booster pump, the pump curve decreases again to case 1, resulting in working point 4.
Again it takes 10-20 seconds to move from point 3 to point 4. At last the pipeline behind the booster pump is filled
with water, resulting in a decrease of the resistance, taking about 400 seconds. The final working point is point 5.
Both Figure 3-10 and Figure 3-11 give an example of the non-stationary effects in a multi-pump/pipeline system.

Page 88 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Pump/Pipeline System.

3.10. Case study.


The aim of this case study is twofold, first it shows events caused by the dynamic behaviour of the system that
cannot be predicted by steady state calculations, second it shows the application of the above theory. A problem
in defining a system and a scenario for the simulation is, that the system can consist of an infinite number of
pump/pipeline combinations, while there also exists an infinite number of solids property/concentration
distributions as a function of time. For this case study, a system is defined consisting of a suction line followed by
three pump/pipeline units. The first pump is a ladder pump, with a speed of 200 rpm, an impeller diameter of 1.5
m and 1050 kW on the axis (see Figure 3-5). The second and the third pump run also at a speed of 200 rpm, have
an impeller diameter of 2.4 m and 3250 kW on the axis (see Figure 3-6). The time constants of all three pumps are
set to 4 seconds. The time constant of the density meter is set to 10 seconds. The suction line starts at 10 m below
water level, has a length of 12 m and a diameter of 0.69 m. The ladder pump is placed 5 m below water level. The
main pump and the booster pump are placed 10 m above water level. The pipeline length between ladder and main
pump is 30 m, between main pump and booster pump 2000 m, as is the length of the discharge line. The pipe
diameters after the ladder pump are 0.61 m. The total simulation lasts about 28 minutes and starts with the pipeline
filled with water. After the pumps are activated, a mixture with a density of 1.6 ton/m 3 enters the suction mouth
for a period of 2 minutes. A sand is used with a d 15 of 0.25 mm, a d50 of 0.50 mm and a d85 of 0.75 mm. This
density block wave moves through the system, subsequently passing the three pumps. For the simulation the
following scenario is used:

00 minutes start of simulation


01 minutes start of ladder pump
04 minutes start of main pump
07 minutes start of booster pump
10 minutes increase mixture density to about 1.6 ton/m3
12 minutes decrease mixture density to water density
12 minutes take sample of density distribution in pipeline
17 minutes take sample of density distribution in pipeline
22 minutes take sample of density distribution in pipeline
28 minutes stop simulation

Copyright © Dr.ir. S.A. Miedema TOC Page 89 of 304


Introduction Dredging Engineering.
P ip e lin e s e c tio n 1
E l a p s e d: 0 0 : 1 2 : 0 2
T im e
D a te
P r o g r a m: P u m p P i p e l i n e S y s t e m
2 .0

1 .8

to n /c u .m
1 .6
: 0 9 :0 8 :3 7
: M a y 1 2 , 1 9 9 6

1 .4

1 .2

1 .0
0 101 202 303 404 505 606 707 808 909 1010
m

P ip e lin e s e c tio n 2
2 .0

1 .8
m /s e c

1 .6

1 .4

1 .2
W in d o w s V 2 . 0 1

1 .0
1010 1111 1212 1313 1414 1515 1616 1717 1818 1919 2020
m

P ip e lin e s e c tio n 3
2 .0

1 .8
m /s e c

1 .6

1 .4

1 .2

1 .0
2020 2121 2222 2323 2424 2525 2626 2727 2828 2929 3030
m

P ip e lin e s e c tio n 4
2 .0

1 .8
c u .m /s e c

1 .6

1 .4

1 .2

1 .0
3030 3131 3232 3333 3434 3535 3636 3737 3838 3939 4040
m

Figure 3-12: The density distribution in the pipeline after 12 minutes.

P ip e lin e s e c tio n 1
E l a p s e d: 0 0 : 1 7 : 4 0
T im e
D a te
P r o g r a m: P u m p P i p e l i n e S y s t e m

2 .0

1 .8
to n /c u .m

1 .6
: 0 9 :1 4 :1 6
: M a y 1 2 , 1 9 9 6

1 .4

1 .2

1 .0
0 101 202 303 404 505 606 707 808 909 1010
m

P ip e lin e s e c tio n 2
2 .0

1 .8
m /s e c

1 .6

1 .4

1 .2
W in d o w s V 2 . 0 1

1 .0
1010 1111 1212 1313 1414 1515 1616 1717 1818 1919 2020
m

P ip e lin e s e c tio n 3
2 .0

1 .8
m /s e c

1 .6

1 .4

1 .2

1 .0
2020 2121 2222 2323 2424 2525 2626 2727 2828 2929 3030
m

P ip e lin e s e c tio n 4
2 .0

1 .8
c u .m /s e c

1 .6

1 .4

1 .2

1 .0
3030 3131 3232 3333 3434 3535 3636 3737 3838 3939 4040
m

Figure 3-13: The density distribution in the pipeline after 17 minutes.

Page 90 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Pump/Pipeline System.
P ip e lin e se ctio n 1
E l a p s e d: 0 0 : 2 1 : 4 8
T im e
D a te
P r o g r a m: P u m p
2 .0

1 .8

to n /c u .m
1 .6
: 0 9 :1 8 :2 3
: M a y 1 2 , 1 9 9 6

1 .4

1 .2

1 .0
0 101 202 303 404 505 606 707 808 909 1010
P ip e lin e

P ip e lin e se ctio n 2
2 .0

1 .8
S y s te m

m /s e c

1 .6

1 .4

1 .2
W in d o w s V 2 . 0 1

1 .0
1010 1111 1212 1313 1414 1515 1616 1717 1818 1919 2020
m

P ip e lin e se ctio n 3
2 .0

1 .8
m /s e c

1 .6

1 .4

1 .2

1 .0
2020 2121 2222 2323 2424 2525 2626 2727 2828 2929 3030
m

P ip e lin e se ctio n 4
2 .0

1 .8
c u .m /s e c

1 .6

1 .4

1 .2

1 .0
3030 3131 3232 3333 3434 3535 3636 3737 3838 3939 4040
m

Figure 3-14: The density distribution in the pipeline after 22 minutes.

Figure 3-12, Figure 3-13 and Figure 3-14 show the density wave at 12, 17 and 22 minutes of simulation time. At
12 minutes the density wave occupies the suction line, the ladder pump and the main pump and part of the pipeline
behind the main pump. At 17 minutes the density wave occupies the last part of the pipeline before the booster
pump, the booster pump and the first part of the discharge line after the booster pump. At 22 minutes the density
wave occupies the middle part of the discharge line. Figure 3-15 shows the line speed, the density, the total power
consumed and the production as a function of time. The line speed, the density and the production are determined
at the inlet of the ladder pump. The density is determined using the mathematical behaviour of a density transducer
with a time constant of 10 seconds. Figure 3-16, Figure 3-17 and Figure 3-18 show the pump speed, power, vacuum
and discharge pressure of the three pumps as a function of time.

As can be seen in Figure 3-15, the line speed increases slower then the pump speed, due to the inertial effect in the
fourth term of equation (3-23). When the density wave passes the ladder and main pump (from 10 to 13 minutes),
the discharge pressure of these pumps increases, resulting in a higher line speed. When the density wave passes
the booster pump (from 16 to 19 minutes) the same occurs for the booster pump.
After about 10 minutes of simulation time, all three pumps are activated and a steady state situation occurs in the
system. Then the mixture density at the suction mouth increases from water density to about 1.6 ton/m 3. First the
resistance in the suction line increases, resulting in a sudden decrease of the ladder pump vacuum and discharge
pressure. When the density wave reaches the ladder pump, the discharge pressure increases, due to the higher
density. When after 2 minutes, the density decreases to the water density, first the resistance in the suction line
decreases, resulting in an increase of the ladder pump vacuum and discharge pressure, followed by a decrease of
the discharge pressure when the clear water reaches the ladder pump (see Figure 3-15). The distance between the
ladder pump and the main pump is 30 m. With an average line speed of 5 m/s, the density wave passes the main
pump 6 seconds after passing the ladder pump. The same phenomena as described for the ladder pump, occur 6
seconds later for the main pump (see Figure 3-17). Due to the increased discharge pressure of ladder and main
pump during the density wave, the line speed will also increase (see Figure 3-15), but because of the inertial effects,
this increase and 2 minutes later decrease is not as steep.

Copyright © Dr.ir. S.A. Miedema TOC Page 91 of 304


Introduction Dredging Engineering.

L in e s p e e d v s tim e

F lo w
M a y 1 4 , 1 9 9 6 , 1 2 :2 7 :4 7 P M
P u m p P ip e lin e S y s te m
1 0 .0

8 .0
T im e S e rie s

m /s e c
6 .0

4 .0

2 .0

0 .0
0 0 :0 0 0 0 :0 3 0 0 :0 6 0 0 :0 9 0 0 :1 2 0 0 :1 5 0 0 :1 8 0 0 :2 1 0 0 :2 4 0 0 :2 7 0 0 :3 0
T im e

D e n s it y v s t im e
W in d o w s V 2 . 0 1

2 .0 0

1 .8 0
to n /m ^ 3

1 .6 0

1 .4 0

1 .2 0

1 .0 0
0 0 :0 0 0 0 :0 3 0 0 :0 6 0 0 :0 9 0 0 :1 2 0 0 :1 5 0 0 :1 8 0 0 :2 1 0 0 :2 4 0 0 :2 7 0 0 :3 0
T im e

T o t a l p o w e r v s t im e
3000

2400

1800
kW

1200

600

0
0 0 :0 0 0 0 :0 3 0 0 :0 6 0 0 :0 9 0 0 :1 2 0 0 :1 5 0 0 :1 8 0 0 :2 1 0 0 :2 4 0 0 :2 7 0 0 :3 0
T im e

P r o d u c tio n v s tim e
4000

3200
m ^ 3 /h o u r

2400

1600

800

0
0 0 :0 0 0 0 :0 3 0 0 :0 6 0 0 :0 9 0 0 :1 2 0 0 :1 5 0 0 :1 8 0 0 :2 1 0 0 :2 4 0 0 :2 7 0 0 :3 0
T im e

Figure 3-15: Line speed, density, total power and situ production as a function of time.

P u m p s p e e d v s t im e
P u m p 1 T im e S e r ie s
M a y 1 4 , 1 9 9 6 , 1 2 :2 8 :4 9 P M
P u m p P ip e lin e S y s te m

400

320

240
rp m

160

80

0
0 0 :0 0 0 0 :0 3 0 0 :0 6 0 0 :0 9 0 0 :1 2 0 0 :1 5 0 0 :1 8 0 0 :2 1 0 0 :2 4 0 0 :2 7 0 0 :3 0
T im e

P u m p p o w e r v s t im e
W in d o w s V 2 . 0 1

1000

800

600
kW

400

200

0
0 0 :0 0 0 0 :0 3 0 0 :0 6 0 0 :0 9 0 0 :1 2 0 0 :1 5 0 0 :1 8 0 0 :2 1 0 0 :2 4 0 0 :2 7 0 0 :3 0
T im e

P u m p v a c u u m v s t im e ( - = v a c u u m , + = p r e s s u r e )
1 0 0 .0

6 0 .0

2 0 .0
kP a

-2 0 .0

-6 0 .0

-1 0 0 .0
0 0 :0 0 0 0 :0 3 0 0 :0 6 0 0 :0 9 0 0 :1 2 0 0 :1 5 0 0 :1 8 0 0 :2 1 0 0 :2 4 0 0 :2 7 0 0 :3 0
T im e

P u m p d is c h a r g e p r e s s u r e v s t im e
300

240

180
kP a

120

60

0
0 0 :0 0 0 0 :0 3 0 0 :0 6 0 0 :0 9 0 0 :1 2 0 0 :1 5 0 0 :1 8 0 0 :2 1 0 0 :2 4 0 0 :2 7 0 0 :3 0
T im e

Figure 3-16: Speed, power, vacuum and discharge pressure of the ladder pump vs. time.

One could say that there is a time delay between the immidiate response of the discharge pressure of the pumps
on changes in the density in the pumps and the response of the line speed on changes in the discharge pressure. At
12 minutes and about 15 seconds, the density wave has left the main pump, but has not yet reached the booster
pump. The head of each pump is determined by the density of water, but the line speed is still determined by the

Page 92 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Pump/Pipeline System.

head resulting from the mixture and thus to high. The resistance in the pipe between main and booster pump is
high because of the mixture, resulting in a decrease of the booster pump vacuum and discharge pressure. As the
line speed decreases, the booster pump vacuum and discharge pressure will stay in a semi-steady state situation.
When the density wave reaches the booster pump, the total head of the booster pump increases, resulting in an
increase of the line speed. This occurs after about 16.5 minutes of simulation time. Since the total head of ladder
and main pump does not change, the booster pump vacuum will have to decrease to pull harder on the mixture in
the pipeline before the booster pump. This results in the occurence of cavitation of the booster pump, limiting the
total head of the booster pump and thus the line speed. The cavitation causes a very instable behaviour of the
booster pump as is shown in Figure 3-18. Since the density wave moves from the suction line to the discharge line,
the booster pump vacuum and discharge pressure both increase when the density wave moves through the booster
pump. After 18.5 minutes the density wave leaves the booster pump. The total head of the booster pump decreases
sharply, while the line speed decreases slowly. The fluid in the pipeline before the booster pump pushes and the
fluid after the booster pump pulls, resulting in a quick increase of the booster pump vacuum and a decrease in the
booster pump discharge pressure. As the line speed decreases, the discharge pressure will increase again. After 23
minutes of simulation time, the density wave starts leaving the pipeline. 2 minutes later the density wave has
complete left the system. Because of the decreasing resistance during this time-span, the line speed will increase
slightly, resulting in a small decrease of the vacuum and discharge pressure of each pump, while the total head
remains constant. The total power will also increase slightly because of this.
To stabilise the line speed to a specific value, flow control can be used. Flow control adjusts the speed of the last
pump, in this case the booster pump. If the line speed is higher then a set point, the booster pump speed is decreased,
if the line speed is lower, the booster pump speed is increased. To determine the correct booster pump speed, the
total head is considered to be a summation of the heads of all of the pumps in the system. The head of the booster
pump is considered to be proportional to the square of the booster pump speed and the total resistance is considered
to be proportional to the square of the line speed, this gives:

2 2
 p l .p .   p m .p .   p b .p .   p l .p .   p m .p .    n   c (3-27)

When the flow control is active, the heads of the ladder pump and the main pump do not change, so for the set
point of the line speed:

2 2
 p l .p .   p m .p .   p b .p .   p l .p .   p m .p .    n f .c .    c f .c . (3-28)

Assuming that the sum of the heads of ladder and main pump equals the head of the booster pump times a factor
 and dividing equation (3-28) by equation (3-27), the following can be derived:

2
 c f .c . 
n f .c .  n    1      (3-29)
 c 

 c f .c .  c 
By substituting:     and using Taylor series approximation, this gives:
 c 

    1      2
1
n f .c .  n  n  (3-30)
2

Equation (3-30) is used to simulate flow control. The same scenario as above is used, except for the flow control
that is activated after 8 minutes of simulation time. The set point for the line speed is set to 5 m/sec. Figure 3-19
and Figure 3-20 show the results of this simulation. As can be seen, the line speed changes rapidly when the density
wave reaches or leaves one of the pumps. In about 15 seconds the flow control has adjusted the line speed to the
set point. Figure 3-20 shows that the occurrence of cavitation is almost surpressed using the flow control. The
booster pump speed tends to slightly oscillate. This is caused by applying several first order systems in series,
resulting in a second or third order system. If the factor  is choosen to high, the system is fast but tends to oscillate.
If this factor is to small, the system responds very slow. In the simulation a value of 2 is used.

Copyright © Dr.ir. S.A. Miedema TOC Page 93 of 304


Introduction Dredging Engineering.
P u m p s p e e d v s t im e

P u m p 2 T im e S e r ie s
M a y 1 4 , 1 9 9 6 , 1 2 :2 9 :2 7 P M
P u m p P ip e lin e S y s te m
400

320

240

rp m
160

80

0
0 0 :0 0 0 0 :0 3 0 0 :0 6 0 0 :0 9 0 0 :1 2 0 0 :1 5 0 0 :1 8 0 0 :2 1 0 0 :2 4 0 0 :2 7 0 0 :3 0
T im e

P u m p p o w e r v s t im e
W in d o w s V 2 . 0 1

3000

2400

1800
kW

1200

600

0
0 0 :0 0 0 0 :0 3 0 0 :0 6 0 0 :0 9 0 0 :1 2 0 0 :1 5 0 0 :1 8 0 0 :2 1 0 0 :2 4 0 0 :2 7 0 0 :3 0
T im e

P u m p v a c u u m v s t im e ( - = v a c u u m , + = p r e s s u r e )
1 0 0 .0

6 0 .0

2 0 .0
kP a

-2 0 .0

-6 0 .0

-1 0 0 .0
0 0 :0 0 0 0 :0 3 0 0 :0 6 0 0 :0 9 0 0 :1 2 0 0 :1 5 0 0 :1 8 0 0 :2 1 0 0 :2 4 0 0 :2 7 0 0 :3 0
T im e

P u m p d is c h a r g e p r e s s u r e v s t im e
700

560

420
kP a

280

140

0
0 0 :0 0 0 0 :0 3 0 0 :0 6 0 0 :0 9 0 0 :1 2 0 0 :1 5 0 0 :1 8 0 0 :2 1 0 0 :2 4 0 0 :2 7 0 0 :3 0
T im e

Figure 3-17: Speed, power, vacuum and discharge pressure of the main pump vs. time.

P u m p s p e e d v s t im e
P u m p 3 T im e S e r ie s
M a y 1 4 , 1 9 9 6 , 1 2 :2 9 :5 2 P M
P u m p P ip e lin e S y s te m

400

320

240
rp m

160

80

0
0 0 :0 0 0 0 :0 3 0 0 :0 6 0 0 :0 9 0 0 :1 2 0 0 :1 5 0 0 :1 8 0 0 :2 1 0 0 :2 4 0 0 :2 7 0 0 :3 0
T im e

P u m p p o w e r v s t im e
W in d o w s V 2 . 0 1

3000

2400

1800
kW

1200

600

0
0 0 :0 0 0 0 :0 3 0 0 :0 6 0 0 :0 9 0 0 :1 2 0 0 :1 5 0 0 :1 8 0 0 :2 1 0 0 :2 4 0 0 :2 7 0 0 :3 0
T im e

P u m p v a c u u m v s t im e ( - = v a c u u m , + = p r e s s u r e )
1 0 0 .0

6 0 .0

2 0 .0
kP a

-2 0 .0

-6 0 .0

-1 0 0 .0
0 0 :0 0 0 0 :0 3 0 0 :0 6 0 0 :0 9 0 0 :1 2 0 0 :1 5 0 0 :1 8 0 0 :2 1 0 0 :2 4 0 0 :2 7 0 0 :3 0
T im e

P u m p d is c h a r g e p r e s s u r e v s t im e
700

560

420
kP a

280

140

0
0 0 :0 0 0 0 :0 3 0 0 :0 6 0 0 :0 9 0 0 :1 2 0 0 :1 5 0 0 :1 8 0 0 :2 1 0 0 :2 4 0 0 :2 7 0 0 :3 0
T im e

Figure 3-18: Speed, power, vacuum and discharge pressure of the booster pump vs. time.

Page 94 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Pump/Pipeline System.
L in e s p e e d v s t im e

F lo w
M a y 1 4 , 1 9 9 6 , 0 9 :1 8 :4 6 P M
P u m p P ip e lin e S y s te m
1 0 .0

8 .0
T im e S e rie s

m /s e c
6 .0

4 .0

2 .0

0 .0
0 0 :0 0 0 0 :0 3 0 0 :0 6 0 0 :0 9 0 0 :1 2 0 0 :1 5 0 0 :1 8 0 0 :2 1 0 0 :2 4 0 0 :2 7 0 0 :3 0
T im e

D e n s it y v s t im e
W in d o w s V 2 . 0 1

2 .0 0

1 .8 0
to n /m ^ 3

1 .6 0

1 .4 0

1 .2 0

1 .0 0
0 0 :0 0 0 0 :0 3 0 0 :0 6 0 0 :0 9 0 0 :1 2 0 0 :1 5 0 0 :1 8 0 0 :2 1 0 0 :2 4 0 0 :2 7 0 0 :3 0
T im e

T o t a l p o w e r v s t im e
3000

2400

1800
kW

1200

600

0
0 0 :0 0 0 0 :0 3 0 0 :0 6 0 0 :0 9 0 0 :1 2 0 0 :1 5 0 0 :1 8 0 0 :2 1 0 0 :2 4 0 0 :2 7 0 0 :3 0
T im e

P r o d u c t io n v s t im e
4000

3200
m ^ 3 /h o u r

2400

1600

800

0
0 0 :0 0 0 0 :0 3 0 0 :0 6 0 0 :0 9 0 0 :1 2 0 0 :1 5 0 0 :1 8 0 0 :2 1 0 0 :2 4 0 0 :2 7 0 0 :3 0
T im e

Figure 3-19: Line speed, density, total power and situ production as a function of time, with flow control.

P u m p s p e e d v s t im e
P u m p 3 T im e S e r ie s
M a y 1 4 , 1 9 9 6 , 0 9 :1 8 :2 1 P M
P u m p P ip e lin e S y s te m

400

320

240
rp m

160

80

0
0 0 :0 0 0 0 :0 3 0 0 :0 6 0 0 :0 9 0 0 :1 2 0 0 :1 5 0 0 :1 8 0 0 :2 1 0 0 :2 4 0 0 :2 7 0 0 :3 0
T im e

P u m p p o w e r v s t im e
W in d o w s V 2 . 0 1

3000

2400

1800
kW

1200

600

0
0 0 :0 0 0 0 :0 3 0 0 :0 6 0 0 :0 9 0 0 :1 2 0 0 :1 5 0 0 :1 8 0 0 :2 1 0 0 :2 4 0 0 :2 7 0 0 :3 0
T im e

P u m p v a c u u m v s t im e ( - = v a c u u m , + = p r e s s u r e )
1 0 0 .0

6 0 .0

2 0 .0
kP a

-2 0 .0

-6 0 .0

-1 0 0 .0
0 0 :0 0 0 0 :0 3 0 0 :0 6 0 0 :0 9 0 0 :1 2 0 0 :1 5 0 0 :1 8 0 0 :2 1 0 0 :2 4 0 0 :2 7 0 0 :3 0
T im e

P u m p d is c h a r g e p r e s s u r e v s t im e
1000

800

600
kP a

400

200

0
0 0 :0 0 0 0 :0 3 0 0 :0 6 0 0 :0 9 0 0 :1 2 0 0 :1 5 0 0 :1 8 0 0 :2 1 0 0 :2 4 0 0 :2 7 0 0 :3 0
T im e

Figure 3-20: Speed, power, vacuum and discharge pressure of the booster pump vs. time, with flow
control.

Copyright © Dr.ir. S.A. Miedema TOC Page 95 of 304


Introduction Dredging Engineering.

3.11. Conclusions and Discussion.


The behaviour of a multi pump/pipeline system is hard to understand. As mentioned before, an infinite number of
system configurations and soil conditions exist. Systems are usually configured, based on steady state calculations,
while the dynamic behaviour is ignored. Combining the steady state approach for pipeline resistance with the
dynamic behaviour of pumps, pump drives and the second law of Newton, the dynamic behaviour can be simulated.
However, a number of assumptions had to be made.

These assumptions are:


1. There is no longitudinal diffusion in the pipeline.
2. The pump drive behaves like a constant torque system.
3. The pipeline resistance is determined using the Durand theory.
4. The centrifugal pump obeys the affinity laws.

Whether these assumptions are valid will be subject of further research. The simulations however show the
occurrence of phenomena that are known in practice.

Multi pump/pipeline systems can be configured in an infinite number of configurations. Phenomena that occur in
one configuration do not have to occur in other configurations. So the configuration to carry out simulations to
examine certain phenomena has to be chosen carefully. The configuration used in this paper is suitable for
simulation of most phenomena. The examples show, that moving from one working point to the next working
point, does not occur instantaneously, but with a time delay, where the time delay depends on the phenomena.

The simulation model used is very well suitable for fully suspended load, but has a deficiency for two phase flow.
The main shortcoming is the fact that suspended load and bed load move through the system at two different
velocities, not being equal to the average line speed.

A second shortcoming is the lack of availability of a good model for the vertical diffusion between the suspended
load and the bed load. This will be subject for further research.

One should consider that mathematical modelling is an attempt to describe reality without having any presumption
of being reality.

Page 96 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Pump/Pipeline System.

3.12. Nomenclature.
C1,2,3,4 Coefficients -
CD Drag coefficient -
Cvt Transport concentration -
Cv Volumetric concentration -
Cvs Spatial concentration -
Cx Drag coefficient -
d Grain/particle diameter m
D Impeller diameter m
Dp Pipe diameter m
Fr Froude number -
g Gravitational constant m/sec2
H Height/elevation m
I Mass moment of inertia tonm3
k Constant -
Kp Proportionality constant kNms/rad
L Length of pipeline m
n Revolutions rpm
p Pressure kPa
P Power kW
3
Q Flow m /sec
r Radius m
Re Reynolds number -
T Torque kNm
u Tangetial velocity m/sec
vt Settling velocity grains m/sec
vls Line speed m/sec
, Coefficients -
 Blade angle rad
 Wall roughness m
 Ratio -
 Efficiency -
 Rotation angle of centrifugal pump rad
 Angular velocity of centrifugal pump rad/sec
 Angular acceleration of centrifugal pump rad/sec2
 Darcy Weisbach friction coefficient -
 Kinematic viscosity 2
m /sec
 Density ton/m3
 Time constant sec
 Friction coefficient -
 Shape factor -

Indices
c Concentration
cr Critical
c.p. Centrifugal pump

Copyright © Dr.ir. S.A. Miedema TOC Page 97 of 304


Introduction Dredging Engineering.

D Design
d Discharge
d.e. Diesel engine
d.f. Dry friction
f Fluid
g Geodetic
gr Grain
g.b. Gear box
h.f. Hydraulic friction
h.i. Hydraulic impact
h.p. Hydraulic power
h.t. Hydraulic transport
i In
m Mixture
m Measured
n Revolutions
o Out
p Proportional
p Pump
p Pipe
q Quarts
s.p. Set point
t. Total
w Water
0 Initial value (boundary condition)
n Number of time step
E Euler
15 % passing
50 % passing
85 % passing

Page 98 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Pump/Pipeline System.

3.13. Notes.

Copyright © Dr.ir. S.A. Miedema TOC Page 99 of 304


Introduction Dredging Engineering.

Page 100 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Pump/Pipeline System.

Copyright © Dr.ir. S.A. Miedema TOC Page 101 of 304


Introduction Dredging Engineering.

Page 102 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Pump/Pipeline System.

Copyright © Dr.ir. S.A. Miedema TOC Page 103 of 304


Introduction Dredging Engineering.

Page 104 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Pressure Losses with Homogeneous Liquid Flow.

Chapter 4: Pressure Losses with Homogeneous Liquid Flow.

4.1. Pipe Wall Shear Stress.


In general objects in a fluid flow experience a resistance proportional to the dynamic pressure of the fluid:

1 2
  l  v ls (4-1)
2

For an object in a fluid flow (like settling particles) the drag force on the object is the dynamic pressure times a
characteristic cross section times a drag coefficient, giving:

1 2
Fd r a g  C D    l  v ls  A (4-2)
2

The drag coefficient normally depends on the Reynolds number of the flow. Now with pipe flow, there is no flow
around an object, but there is flow inside the pipe. The basic principles however remain the same, giving for the
wall shear stress:

1 2
w  f    l  v ls (4-3)
2

The proportionality coefficient f is the so called Fanning friction factor, named after John Thomas Fanning (1837-
1911). The friction force or drag force on a pipe with diameter Dp and length ΔL is now:

1 2
Fd r a g   w  A  f    l  v ls    D p   L (4-4)
2

The pressure difference over the pipe with diameter Dp and length L is the drag force divided by the pipe cross
section Ap:

1 2
f    l  v ls    D p   L
Fd r a g
2 L 2
p    2f    l  v ls (4-5)
A  2 Dp
p D p
4

The notation using the Darcy friction factor also called the Darcy Weisbach friction factor or the Moody friction
factor is more convenient here for using the dynamic pressure, giving:

L 1 2
pl   l     l  v ls (4-6)
Dp 2

Note that the Darcy Weisbach friction factor is 4 times the Fanning friction factor. In terms of the shear stress this
gives:

l 1 2 l 2
w     l  v ls    l  v ls (4-7)
4 2 8

The hydraulic gradient iw (for water) or il (for a liquid in general) is:

2
pl  l  v ls (4-8)
il  iw  
l  g  L 2 g Dp

In this book the Darcy Weisbach friction factor is used.

Copyright © Dr.ir. S.A. Miedema TOC Page 105 of 304


Introduction Dredging Engineering.

4.2. The Darcy-Weisbach Friction Factor.


The value of the wall friction factor l depends on the Reynolds number:

v ls  D p  l  v ls  D p
Re   (4-9)
l l

For laminar flow (Re<2320) the value of l can be determined according to Poiseuille:

64
l  (4-10)
Re

For turbulent flow (Re>2320) the value of l depends not only on the Reynolds number but also on the relative
roughness of the pipe /Dp. A general implicit equation for l is the Colebrook-White (1937) equation:

1
 l 
2
  2 .5 1 0 .2 7     (4-11)
 2  lo g 1 0    
  R e  Dp  
  l  

For very smooth pipes the value of the relative roughness /Dp is almost zero, resulting in the Prandl & von Karman
equation:

1
 l 
2
  2 .5 1   (4-12)
 2  lo g 1 0   
  R e   
  l  

At very high Reynolds numbers the value of 2.51/(Rel) is almost zero, resulting in the Nikuradse (1933)
equation:

1 5 .3
l  
2 2
  0 .2 7       0 .2 7     (4-13)
 2  lo g 1 0    2  ln  
  Dp    Dp 
     

Because equations (4-11) and (4-12) are implicit, for smooth pipes approximation equations can be used. For a
Reynolds number between 2320 and 105 the Blasius equation gives a good approximation:

0 .2 5
 1 
 l  0 .3 1 6 4    (4-14)
 Re 

For a Reynolds number in the range of 105 to 108 the Nikuradse (1933) equation gives a good approximation:

0 .2 2 1
 l  0 .0 0 3 2  (4-15)
0 .2 3 7
Re

Figure 4-1 gives the so called Moody (1944) diagram, in this case based on the Swamee Jain (1976) equation.

Page 106 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Pressure Losses with Homogeneous Liquid Flow.

The legend shows the relative roughness. Moody Diagram


Darcy Weisbach friction coefficient.
Moody diagram for the determination of the
0.08 0.08

0.07 0.07

0.06 0.06

0.05 0.05

Labda
Labda

0.04 0.04

0.03 0.03

0.02 0.02

0.01 0.01

0.00 0.00
2 3 4 5 6 7 8
10 10 10 10 10 10 10
Re
Laminar Smooth 0.000003 0.00001 0.0001 0.0003 0.001

0.003 0.01 0.02 0.03 0.04 0.05

Figure 4-1: The Moody diagram determined with the Swamee Jain equation.

Over the whole range of Reynolds numbers above 2320 the Swamee Jain (1976) equation gives a good
approximation:

1 .3 2 5 0 .2 5
l  
2 2
   5 .7 5      5 .7 5   (4-16)
 ln      lo g 1 0    
  3 .7  D Re
0 .9     3 .7  D Re
0 .9  
  p     p  

4.3. The Equivalent Liquid Model.


Assuming that the pressure losses in a pipe are proportional to the kinetic energy of the eddies and the kinetic
energy of the eddies is proportional to the mixture density and the line speed and assuming that there are no losses
due to sliding friction or collisions, the pressure losses can be determined by:

L 1 2
pm   l     m  v ls (4-17)
Dp 2

The hydraulic gradient im (for mixture) is now:

2
pm m  l  v ls m (4-18)
im      il
l  g  L l 2 g Dp l

The above assumptions are valid as long as the particles are small enough to be considered part of the eddies. So
for larger particles this may not be true anymore.

Copyright © Dr.ir. S.A. Miedema TOC Page 107 of 304


Introduction Dredging Engineering.

4.4. Approximation of the Darcy-Weisbach Friction Factor.


It is obvious that the Darcy-Weisbach friction factor λl depends on the pipe diameter Dp and the line speed vls. This
may be confused with a direct influence of the pipe diameter Dp and the line speed vls. So it is interesting to see
how the Darcy-Weisbach friction factor λl depends on the pipe diameter Dp and the line speed vls. Figure 4-2 shows
the Darcy-Weisbach friction factor for smooth pipes as a function of the line speed vls at a number of pipe
diameters, while Figure 4-3 shows the Darcy-Weisbach friction factor as a function of the pipe diameter Dp at a
number of line speeds. In both figures, the Darcy-Weisbach friction factor can be well approximated by a power
function

1 2
 l     v ls   Dp  (4-19)

With:

 0 .0 8 9  0 .0 8 8
  0 .0 1 2 1 6 and 
 1   0 .1 5 3 7  D p  and  2   0 .2 0 1 3   v ls  (4-20)

For laboratory conditions both powers are close to -0.18, while for real life conditions with higher line speeds and
much larger pipe diameters this results in a power for the line speed of about α1=-0.155 and for the pipe diameter
of about α2=-0.175. This should be considered when analyzing the models for heterogeneous transport.

4.5. The Friction Velocity or Shear Velocity u*.


The term friction velocity comes from the fact that √(τw/ρl) has the same unit as velocity and it has something to
do with the wall friction force. The wall shear stress τw is often represented by friction velocity u*, defined by:

w l
u*    v ls (4-21)
l 8

4.6. The Thickness of the Viscous Sub Layer δv.


Very close to the pipe wall the flow is laminar in the so called viscous sub layer. The thickness of the viscous sub
layer is defined as:

l
 v  1 1 .6  (4-22)
u*

4.7. The Smallest Eddies.


The ratio between the largest eddies and the smallest eddies in turbulent pipe flow is of the magnitude of the
Reynolds number to the power of ¾. Assuming that the largest eddies are of the magnitude of the pipe diameter,
then this gives for the diameter of the smallest eddies:

1/4 3/4
Dp Dp l
de   (4-23)
3/4 3/4
Re v ls

Using the Blasius equation for the Darcy Weisbach friction factor, this gives for the ratio between the diameter of
the smallest eddies to the thickness of the viscous sub layer:

de 1/8
 0 .0 1 7  R e (4-24)
v

For Reynolds numbers ranging from 100,000 for small pipe diameters to 10,000,000 for large pipe diameters this
gives a ratio of 0.072 to 0.127, so about 10%.

Page 108 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Pressure Losses with Homogeneous Liquid Flow.

Darcy-Weisbach friction factor λ vs. Line speed vls


0.030
Dp=0.0254 m

Dp=0.0508 m

0.025 Dp=0.1016 m

Dp=0.1524 m
Darcy-Weisbach friction factor λ (-)

Dp=0.2032 m
0.020 Dp=0.2540 m

Dp=0.3000 m

Dp=0.4000 m
0.015
Dp=0.5000 m

Dp=0.6000 m

0.010 Dp=0.7000 m

Dp=0.8000 m

Dp=0.9000 m
0.005 Dp=1.0000 m

Dp=1.1000 m

Dp=1.2000 m
0.000
1 2 3 4 5 6 7
Line speed vls (m/sec)
© S.A.M.

Figure 4-2: The Darcy-Weisbach friction factor λl for smooth pipes as a function of the line speed v ls.

Darcy-Weisbach friction factor λ vs. Pipe diameter Dp


0.030

vls=1.00 m/sec

0.025
vls=2.00 m/sec
Darcy-Weisbach friction factor λ (-)

0.020
vls=3.00 m/sec

0.015 vls=4.00 m/sec

vls=5.00 m/sec
0.010

vls=6.00 m/sec
0.005

vls=7.00 m/sec

0.000
0 0.2 0.4 0.6 0.8 1 1.2
Pipe diameter Dp (m)
© S.A.M.

Figure 4-3: The Darcy-Weisbach friction factor λl for smooth pipes as a function of the pipe diameter Dp.

Copyright © Dr.ir. S.A. Miedema TOC Page 109 of 304


Introduction Dredging Engineering.

4.8. The Apparent Viscosity.


Einstein (1905) published an analysis for the viscosity of dilute suspensions. The result of this analysis is an
equation giving the relation between the apparent dynamic viscosity and the volumetric concentration of the solids.
The concentrations however are limited to low concentrations.

m
r   1  2 .5  C vs (4-25)
l

Thomas (1965) collected data regarding the relative viscosity from 16 sources. The particle materials included
polystyrene, rubber latex, glass and methyl methacrylate. The results are shown in Figure 4-4. In all studies, either
the density of the suspending medium was adjusted or the viscosity of the suspending medium was sufficiently
large that settling was unimportant. Examination of the experimental procedure used in these studies shows no
basis for eliminating any of the data because of faulty technique; consequently, there must be at least one additional
parameter that has not been accounted for. One parameter of importance is the absolute value of the particle
diameter. For particles with diameters less than 1 to 10 microns, colloid-chemical forces become important causing
non-Newtonian flow behavior. The result is a relative viscosity which increases as particle size is decreased, but
which decreases to a limiting value as the shear rate is increased. For particles larger than 1 to 10 microns, the
inertial effects due to the restoration of particle rotation after collision result in an additional energy dissipation
and consequent increase in relative viscosity with increasing particle diameter.
In flow through capillary tubes, the increase in viscosity observed with large particle size suspensions is opposed
by a decrease in viscosity caused by a tendency for particles to migrate toward the center of the tube as the particle
diameter is increased. Examination of the data from which Figure 4-4 was prepared showed that in several cases
the tests covered a sufficient range of shear rates or partied sizes that it was possible to extrapolate to conditions
where particle size effects were negligible. For particles less than 1 micron diameter, the limiting value of the
relative viscosity was obtained as the intercept of either a linear plot of 1/d versus μm/μl or a linear plot of 1/(du/dr)
versus μm/μl. For particles larger than 1 to 10 microns, the limiting value of the relative viscosity was obtained as
the intercept of a linear plot of d versus μm/μl• In the event that large particle size data were also available as a
function of shear rate, the reduced particle size data were further corrected by plotting against 1/(du/ dr). Treatment
of the suitable data in this manner gave a unique curve for which the maximum deviation was reduced from three-
to six fold over that shown in Figure 4-4, that is, to ± 7 % at Cvs=0.2 and to ±13 % at Cvs=0.5, as is show in Figure
4-5.

Based on this Thomas (1965) derived an equation to determine the relative dynamic viscosity as a function of the
concentration Cvs of the particles in the mixture.

m 2 1 6 .6  C v s
r   1  2 .5  C vs  1 0 .0 5  C vs  0 .0 0 2 7 3  e (4-26)
l

The Thomas (1965) equation can be used for pseudo homogeneous flow of small particles.

Figure 4-5 shows that the first two terms are valid to a volumetric concentration of about 6%. Adding the 3 rd term
extends the validity to a volumetric concentration of about 25%. Adding the 4 th term extends the validity to a
volumetric concentration of 60%, which covers the whole range of concentrations important in dredging
applications.

Figure 4-6 shows experiments of Boothroyde et al. (1979) with Markham fines (light solids, high concentration)
without using the Thomas (1965) viscosity. Figure 4-7 shows these experiments using the Thomas (1965)
viscosity.

Figure 4-8 shows experiments of Thomas (1976) with iron ore (very heavy solids, medium concentration) without
using the Thomas (1965) viscosity. Figure 4-9 shows these experiments using the Thomas (1965) viscosity.

In both cases the data points are above the ELM curves if the normal liquid viscosity is used. Using the Thomas
(1965) viscosity correction places the data points very close to the ELM curves. Applying the Thomas (1965)
viscosity gives a good result for the fines, as long as they behave like a Newtonian fluid.

Page 110 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Pressure Losses with Homogeneous Liquid Flow.

Collected Relative Viscosity Data, From 16 Sources

100
Relative Viscosity νm/νl (-)

Experiments
Fit Line 4 Terms
10 Upper Limit
Lower Limit

1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Volume Fraction Solids Cv (-)
© S.A.M.

Figure 4-4: Collected relative viscosity data from 16 sources by Thomas (1965).

Collected Relative Viscosity Data, Reduced


100
Relative Viscosity νm/νfl (-)

Experiments
10
Fit Line 2 Terms
Fit Line 3 Terms
Fit Line 4 Terms

1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Volume Fraction Solids Cv (-)
© S.A.M.

Figure 4-5: Collected relative viscosity data from 16 sources by Thomas (1965), reduced.

The limiting particle diameter for particles influencing the viscosity can be determined based on the Stokes
number. A Stokes number of Stk=0.03 gives a good first approximation. The velocity in the denominator can be
replaced by 7.5·Dp0.4 as a first estimate of the LDV near operational conditions.

S tk  9   l   l  D p S tk  9   l   l  D p
d=  (4-27)
 s  v ls 0 .4
 s  7 .5  D p

Copyright © Dr.ir. S.A. Miedema TOC Page 111 of 304


Introduction Dredging Engineering.

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il


Sliding Bed Cvs=c.

Equivalent Liquid Model


1.000
Relative excess hydraulic gradient Erhg (-)

Homogeneous Flow
Cvs=Cvt=c.
Resulting Erhg curve
Cvs=c.
Resulting Erhg curve
Cvt=c.
0.100
Limit Deposit Velocity

Ratio Potential/Kinetic
Energy
Heterogeneous Flow
with Near Wall Lift
Homogeneous Flow
0.010 Mobilized
Cv=0.481

Cv=0.450

Cv=0.400

0.001
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)
© S.A.M. Dp=0.2000 m, d=0.200 mm, Rsd=0.463, Cv=0.450, μsf=0.416

Figure 4-6: Markham fines Boothroyde et al. (1979), without Thomas (1965) viscosity.

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il


Sliding Bed Cvs=c.

Equivalent Liquid
1.000
Model
Relative excess hydraulic gradient Erhg (-)

Homogeneous Flow
Cvs=Cvt=c.

Uniform Sand Cvs=c.

Uniform Sand Cvt=c.


0.100

Limit Deposit Velocity

Graded Sand Cvs=c.

Graded Sand Cvt=c.


0.010
Cv=0.481

Cv=0.450

Cv=0.400

0.001
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)
© S.A.M. Dp=0.2000 m, d=0.200 mm, Rsd=0.463, Cv=0.450, μsf=0.416

Figure 4-7: Markham fines Boothroyde et al. (1979), with Thomas (1965) viscosity.

Figure 4-6 shows experimental data versus the DHLLDV Framework for uniform particles with the pure liquid
viscosity. The data do not match the curve, but are much higher. Figure 4-7. Shows the experimental data versus
the DHLLDV Framework for graded particles according to Boothroyde et al. (1979) and full Thomas (1965)
viscosity. Now the data match the curve for graded particles.

Page 112 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Pressure Losses with Homogeneous Liquid Flow.

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il


Sliding Bed Cvs=c.

1.000 Equivalent Liquid Model


Relative excess hydraulic gradient Erhg (-)

Homogeneous Flow
Cvs=Cvt=c.

Resulting Erhg curve


Cvs=c.
0.100 Resulting Erhg curve
Cvt=c.

Limit Deposit Velocity

Ratio Potential/Kinetic
Energy
0.010
Heterogeneous Flow
with Near Wall Lift

Homogeneous Flow
Mobilized

Cv=0.240

0.001
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)
© S.A.M. Dp=0.1585 m, d=0.040 mm, Rsd=3.976, Cv=0.240, μsf=0.416

Figure 4-8: Iron ore Thomas (1976), without Thomas (1965) viscosity.

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il


Sliding Bed Cvs=c.

1.000
Equivalent Liquid
Relative excess hydraulic gradient Erhg (-)

Model

Homogeneous Flow
Cvs=Cvt=c.

Uniform Sand Cvs=c.


0.100

Uniform Sand Cvt=c.

Limit Deposit Velocity

0.010 Graded Sand Cvs=c.

Graded Sand Cvt=c.

Cv=0.240

0.001
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)
© S.A.M. Dp=0.1585 m, d=0.040 mm, Rsd=3.976, Cv=0.240, μsf=0.416

Figure 4-9: Iron ore Thomas (1976), with Thomas (1965) viscosity.

Figure 4-8 shows the experimental data versus the DHLLDV Framework. Figure 4-9 shows these experiments
using the Thomas (1965) viscosity based on the particle size distribution mentioned by Thomas (1976). The data
points now match the DHLLDV Framework for graded particles and adjusted viscosity

Copyright © Dr.ir. S.A. Miedema TOC Page 113 of 304


Introduction Dredging Engineering.

4.9. Nomenclature.
Aobj Cross section of object perpendicular to velocity direction m2
Apw Pipe wall surface m2
Ap Pipe cross section m2
CD Drag coefficient -
Cvs Spatial volumetric concentration -
d Particle diameter m
de Diameter smallest eddy m
Dp Pipe diameter m
f Fanning friction factor -
Fdrag Drag force on object N
g Gravitational constant 9.81 m/sec2 m/s2
i, il, iw Hydraulic gradient liquid m.w.c./m
im Hydraulic gradient mixture m.w.c./m
ΔL Length of pipeline m
Δpl Pressure difference over length ΔL liquid kPa
Δpm Pressure difference over length ΔL pseudo homogeneous mixture kPa
Re Reynolds number of pipe flow -
Stk Stokes number -
u* Friction velocity m/s
vls Line speed m/s
α Proportionality constant -
α1 Power of line speed -
α2 Power of pipe diameter -
δv Thickness viscous sub-layer m
ε Roughness of pipe wall m
λl Darcy-Weisbach friction factor liquid to wall -
ρl Liquid density ton/m3
ρm Mixture density ton/m3
μl Dynamic viscosity liquid Pa·s
μm Dynamic viscosity pseudo homogeneous mixture Pa·s
μr Relative dynamic viscosity -
νl Kinematic viscosity liquid m2/s
νm Kinematic viscosity pseudo homogeneous mixture m2/s
νr Relative kinematic viscosity -
τw Wall shear stress Pa

Page 114 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Pressure Losses with Homogeneous Liquid Flow.

4.10. Notes.

Copyright © Dr.ir. S.A. Miedema TOC Page 115 of 304


Introduction Dredging Engineering.

Page 116 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Pressure Losses with Homogeneous Liquid Flow.

Copyright © Dr.ir. S.A. Miedema TOC Page 117 of 304


Introduction Dredging Engineering.

Page 118 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Pressure Losses with Homogeneous Liquid Flow.

Copyright © Dr.ir. S.A. Miedema TOC Page 119 of 304


Introduction Dredging Engineering.

Page 120 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Delft Head Loss & Limit Deposit Velocity Framework.

Chapter 5: The Delft Head Loss & Limit Deposit Velocity Framework.

5.1. Introduction.
5.1.1. Considerations.
In the last decades many head loss models for slurry transport have been developed. Not just for the dredging
industry but also for coal and phosphate transport and in the chemical industries. Some models are based on the
phenomena occurring combined with dimensionless parameters, resulting in semi-empirical equations (Durand &
Condolios (1952), Gibert (1960), Worster & Denny (1955), Jufin Lopatin (1966), Zandi & Govatos (1967),
Fuhrboter (1961)), while others are based on physics with 2 and 3 layer models (Newitt et al. (1955), Wasp et al.
(1977), Doron & Barnea (1987), Wilson (1979), the SRC model (1991) and Matousek (2009)). The physical
models are based on stationary transport in time and space, while the semi-empirical models may incorporate non-
stationary or dynamical processes. An analysis of these models and of data collected from numerous publications
for particles with densities ranging from 1.14 ton/m3 to 3.65 ton/m3, particle diameters ranging from 0.005 mm up
to 45 mm, concentrations up to 45% and pipe diameters from 0.0254 m up to 0.9 m has led to an overall model of
head losses in slurry transport, a sort of Framework. The Framework is based on 5 main flow regimes determining
the source of energy losses, the fixed or stationary bed regime, the sliding bed regime, the heterogeneous flow
regime, the homogeneous flow regime and the sliding flow regime.. One can distinguish viscous friction losses,
dry friction losses, potential energy losses, kinetic energy losses, Magnus lift work, turbulent lift work and
turbulent eddy work. The losses do not have to occur at the same time. Usually one or two will be dominant
depending on the flow regime.

Although sophisticated 2 and 3 layer models exist for slurry flow (here the flow of sand/gravel water mixtures),
the main Dutch and Belgium dredging companies still use modified Durand & Condolios (1952) and Fuhrboter
(1961) models, while the main dredging companies in the USA and Canada use a modified Wilson et al. (1992)
model for heterogeneous transport and sliding bed transport or the SRC model. When asked why these companies
don’t use the more sophisticated models, they answer that they require models that match their inputs and they feel
that the 2 and 3 layer models are still in an experimental phase, although these models give more insight in the
physics. Usually the companies require a model based on the particle size distribution or d50, the pipe diameter Dp,
the line speed vls, the relative submerged density Rsd and the temperature (the viscosity of the carrier liquid νl).
Parameters like the bed associated hydraulic radius are not known in advance and thus not suitable. Usually the
dredging companies operate at high line speeds above the Limit Deposit Velocity (LDV) in the heterogeneous or
homogeneous regime. This implies that the bed has dissolved and 2 and 3 layer models are not applicable anyway.

Still there is a need for improvement, since the existing models give reasonably good predictions for small diameter
pipes, but not for large diameter pipes as used in dredging. Recent projects require line lengths up to 35 km with
5 to 6 booster pumps and large diameter pipes. Choosing the number of booster pumps and the location of the
booster pumps depends on the head losses. However it should be considered that the slurry transport process is not
stationary. Densities may vary from a water density of 1 ton/m3 to densities of 1.6 ton/m3 and particle size
distributions will change over time. This results in a dynamic process where pumps, pump drives and slurry
transport interact. The fundamental 2 and 3 layer models require a stationary approach, while the more empirical
equations may take the dynamic effects as time and place averaged effects into account. The question is whether
a semi empirical approach is possible, covering the whole range of pipe diameters and giving the empirical
equations a more physical background, but still using the parameters available to the dredging industry.

Transporting sand with water through a pipeline, in general, results in an increase of the pressure required
compared with pumping water or pure liquid. Since pressure times flow equals power and power times time equals
energy, this can also be interpreted as an increase of the energy required to pump the solids. Energy or work also
equals force times distance or stress times volume. The fact that more power is required to pump a solid-liquid
mixture compared with just pumping the liquid implies that there are additional energy losses and energy
dissipation when pumping the solids. In order to go into detail to the model developed, first the different types of
energy dissipation due to the solids effect are discussed.

It is clear that the flow regimes and the magnitude of the relative excess hydraulic gradient depends strongly on
the pipe diameter and the particle diameter. In the large pipe a sliding bed will never occur in the constant spatial
volumetric concentration case. In the small pipe however it will for particles larger than 0.5 mm. In the small pipe,
the larger particles exceed the ratio d/Dp>0.015 as set by Wilson et al. (1997), resulting in almost 100% stratified

Copyright © Dr.ir. S.A. Miedema TOC Page 121 of 304


Introduction Dredging Engineering.

flow, here considered to be the sliding flow regime. In the large pipe this criterion is never met, except when
pumping large gravel, cobbles or boulder sized pieces such as cut rock or clayballs.
.
Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il
Equivalent Liquid Model

Homogeneous
1.000
Relative excess hydraulic gradient Erhg (-)

Sliding Bed Cvs=c.

d=0.10 mm, Cvs=c.

d=0.20 mm, Cvs=c.

0.100 d=0.30 mm, Cvs=c.

d=0.50 mm, Cvs=c.

d=0.75 mm, Cvs=c.

d=1.00 mm, Cvs=c.

0.010
d=1.50 mm, Cvs=c.

d=3.00 mm, Cvs=c.

d=10.0 mm, Cvs=c.

Limit Deposit Velocity


0.001
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)

© S.A.M. Dp=0.1524 m, Rsd=1.585, Cvs=0.175, μsf=0.416

Figure 5-1: The relative excess hydraulic gradient as a function of the hydraulic gradient,
constant Cvs and Dp=0.1524 m.

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il


Equivalent Liquid Model

Homogeneous
1.000
Relative excess hydraulic gradient Erhg (-)

Sliding Bed Cvs=c.

d=0.10 mm, Cvt=c.

d=0.20 mm, Cvt=c.

0.100 d=0.30 mm, Cvt=c.

d=0.50 mm, Cvt=c.

d=0.75 mm, Cvt=c.

d=1.00 mm, Cvt=c.

0.010
d=1.50 mm, Cvt=c.

d=3.00 mm, Cvt=c.

d=10.0 mm, Cvt=c.

Limit Deposit Velocity


0.001
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)

© S.A.M. Dp=0.1524 m, Rsd=1.585, Cvt=0.175, μsf=0.416

Figure 5-2: The relative excess hydraulic gradient as a function of the hydraulic gradient,
constant Cvt and Dp=0.1524 m.

Page 122 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Delft Head Loss & Limit Deposit Velocity Framework.

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il


Equivalent Liquid Model

Homogeneous
1.000
Relative excess hydraulic gradient Erhg (-)

Sliding Bed Cvs=c.

d=0.10 mm, Cvs=c.

d=0.20 mm, Cvs=c.

0.100 d=0.30 mm, Cvs=c.

d=0.50 mm, Cvs=c.

d=0.75 mm, Cvs=c.

d=1.00 mm, Cvs=c.

0.010
d=1.50 mm, Cvs=c.

d=3.00 mm, Cvs=c.

d=10.0 mm, Cvs=c.

Limit Deposit Velocity


0.001
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)

© S.A.M. Dp=1.0000 m, Rsd=1.585, Cvs=0.175, μsf=0.416

Figure 5-3: The relative excess hydraulic gradient as a function of the hydraulic gradient,
constant Cvs and Dp=1 m.

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il


Equivalent Liquid Model

Homogeneous
1.000
Relative excess hydraulic gradient Erhg (-)

Sliding Bed Cvs=c.

d=0.10 mm, Cvt=c.

d=0.20 mm, Cvt=c.

0.100 d=0.30 mm, Cvt=c.

d=0.50 mm, Cvt=c.

d=0.75 mm, Cvt=c.

d=1.00 mm, Cvt=c.

0.010
d=1.50 mm, Cvt=c.

d=3.00 mm, Cvt=c.

d=10.0 mm, Cvt=c.

Limit Deposit Velocity


0.001
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)

© S.A.M. Dp=1.0000 m, Rsd=1.585, Cvt=0.175, μsf=0.416

Figure 5-4: The relative excess hydraulic gradient as a function of the hydraulic gradient,
constant Cvt and Dp=1 m.

Figure 5-1, Figure 5-2, Figure 5-3 and Figure 5-4 show the results of the energy approach for 9 sands ranging
from d=0.1 mm up to d=10 mm in pipes with diameters of Dp=0.1524 m and Dp=1 m. For each pipe diameter the
constant spatial volumetric concentration curves and the constant delivered volumetric concentration curves are
shown.

Copyright © Dr.ir. S.A. Miedema TOC Page 123 of 304


Introduction Dredging Engineering.

On the Erhg graph the regimes are clearly distinguishable: the fixed bed is an upward-sloping line on the left, the
sliding bed regime is the flat (horizontal) line at about 0.4 (the sliding friction factor), the heterogeneous regime is
the downward-sloping line in the middle, and the pseudo-homogeneous regime is upward-sloping on the left. For
large particles the sliding flow regime is almost flat (similar to sliding bed), in the flow region of the heterogeneous
regime. The figures above clearly show how the available regimes change with pipe diameter.

So head losses from experiments in pipes of 0.1524 m can hardly be translated into head losses for a pipe of 1 m
as often used in dredging. The physical processes are different. Small pipe sliding bed versus large pipe no sliding
bed and small pipe sliding flow versus large pipe no sliding flow. In fact the smaller the pipe diameter, the higher
the probability of the occurrence of a sliding bed and sliding flow and the larger the pipe diameter, the lower the
probability of the occurrence of a sliding bed and sliding flow. Only if the physical processes involved are similar,
is scaling possible.

This explains why many equations and models from literature give good results for small pipe diameters, but
deviate for large diameter pipes. The way energy is dissipated in small diameter pipes is often different from the
way it is dissipated in large diameter pipes at operational line speeds. It also explains while a lot of research is
focused on 2 and 3 layer transport with a sliding bed, which often occurs in small pipes, but much less in large
pipes.

5.1.2. Energy Dissipation.


When a liquid is transported through a pipeline, energy is dissipated by viscous friction and by turbulence
(assuming high Reynolds numbers). When solids are added, there will also be energy dissipation in the form of
potential losses, kinetic losses and possibly friction losses and losses due to Magnus and turbulent lift work and
turbulence in general.
 Potential energy losses. In turbulent flow, because the solids are under the influence of gravity and the
turbulence has to keep them floating. The potential energy losses will depend on the terminal settling velocity
and be influenced by hindered settling. Since the settling process does not depend on the line speed, at a higher
line speed the energy dissipation per unit of time will not change. This implies that the energy dissipation per
unit of pipeline length is reversely proportional with the line speed. So at high line speeds the influence of the
potential energy losses will diminish.
 Kinetic energy losses. During transport, because the particles interact with the wall, with each other and with
the turbulent eddies and in all cases they lose part of their kinetic energy. With the kinetic energy losses one
may expect that the number of interactions is more or less constant in time, so at higher line speeds the number
of interactions per unit of line length will decrease reversely proportional with the line speed, resulting in a
decrease of the excess pressure due to the solids. At higher line speeds however the momentum of the particles
also increases and it is more difficult to change the direction of the particles. This might decrease the number
of interactions with the wall per unit of time. The total losses will be reversely proportional with the line speed
to a power higher than 1, let’s say a proportionality with a power between -1 and -2. The proportionality
depends on the physical properties and the grading of the solids. Although near wall lift will exist at low line
speeds, it is negligible until a certain line speed where the lift force is strong enough to keep the solids away
from the wall. At this line speed there are no more interactions with the wall and the excess pressure due to
interactions collapses. At about the same line speed the lift forces start driving the solids to the center of the
pipe resulting in a more homogeneous flow. The pure heterogeneous regime stops abruptly, because there are
no more interactions with the wall, and the pseudo homogeneous regime starts, based on the work of lift forces
and turbulence. The transition line speed depends on the particle and the pipe diameter. So the sudden regime
change as described will only occur in uniform or very narrow graded sands.
 Sliding and rolling friction. Sliding and rolling friction occur if there is a sliding or moving bed. Forces are
transmitted directly between particles and the internal and external friction coefficients determine the friction
forces. These coefficients are dependent on the type of solids and the particle size distribution.
 Magnus lift work. When the thickness of the viscous sub-layer is bigger than the particle diameter, particles
with rotation due to interactions with the wall will be subjected to Magnus lift forces. This will only occur for
the combination of a low line speed and small particles. The Magnus lift forces will carry out work if they
actually lift the particles, contributing to the energy losses. When the line speed increases, the thickness of the
viscous sub-layer decreases and the particles do not fit in the viscous sub-layer anymore. The Magnus lift
work will diminish when the size of the particles is bigger than the layer thickness. At a higher line speed, the
turbulent lift and turbulent eddies will take over.
 Turbulent lift and eddy work. At high line speeds the turbulent lift and turbulent eddies becomes important.
Since lift force times the distance over which it acts equals the work carried out, this will also result in energy
losses. Since the lift force increases with the velocity gradient near the wall, the losses due to the lift force

Page 124 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Delft Head Loss & Limit Deposit Velocity Framework.

will increase with the line speed. At relatively low line speeds most solids will be transported in the bottom
part of the pipeline, resulting in an asymmetrical concentration profile, matching heterogeneous flow. This
results in an opposite asymmetrical velocity profile, with the highest flow at the top of the pipeline. Below a
certain line speed the lift force on a particle is smaller than the weight of the particle and the lift force will not
carry out any work. But above this transition velocity suddenly the particles will be lifted. The lift forces are
dependent on the velocity gradient and thus will appear at the full circumference of the pipe, but they will first
start pushing the solids upwards from the bottom and thus start to create a more symmetrical concentration
and velocity profile. With increasing line speed the concentration and velocity profile will get closer to the
symmetrical profiles, matching pseudo homogeneous transport.

Resuming it can be stated that the potential and kinetic losses decrease with an increasing line speed with a power
of the line speed between -1 and -2, while the losses due to near wall lift forces increase with an increasing line
speed, until the pseudo homogeneous regime is reached. For each combination of particle and pipe diameter, there
exists a transition line speed. Below this line speed kinetic losses dominate the excess pressure; above this line
speed the work carried out by turbulent lift and eddy forces dominates the excess pressure. For uniform sand,
kinetic losses and work carried out by lift forces will not occur at the same line speed. For graded sands a transition
region, with respect to the line speed, will occur, the size of which depending on the grading. In the case where
the particles are much smaller than the thickness of the viscous sub layer, theoretically there is Magnus lift if the
particles are rotating. One may expect that the excess pressure due to the solids will continue decreasing with
increasing line speed. In this case the excess pressure will reach zero asymptotically and there is no solids effect
at very high line speeds. It is obvious that the collapse of the interactions with the wall, resulting in a collapse of
the kinetic losses, due to the lift force, will happen at about the same line speed where the work of the lift forces
starts increasing. This is the transition line speed between heterogeneous and pseudo homogeneous transport. It is
not possible that the collapse of the kinetic losses appears at a line speed higher than the line speed where the work
carried out by the lift forces starts, for uniform sands. It might be possible that this collapse appears at a slightly
lower line speed, resulting in a collapse of the excess pressure, but at higher line speeds this will increase again
because of the work of the lift forces.

Wilson et al. (1997) introduced the Stratification Ratio R, which in fact equals the relative excess hydraulic
gradient Erhg. The higher the Stratification Ratio, the more asymmetrical the concentration and the velocity profile
in the pipe. With increasing line speed, the Stratification Ratio decreases with power of 0.25-1.7, depending on the
grading of the sand. However, once the transition line speed between heterogeneous and homogeneous transport
is passed, the relative excess hydraulic gradient will increase again, while the stratification decreases. The term
Stratification Ratio corresponds with the heterogeneous transport, with potential and kinetic losses, but not with
the pseudo homogeneous transport with losses due to lift work. Therefore a new term is introduced, the Slip
Relative Squared or Srs, which is the ratio between the slip velocity and the terminal settling velocity squared.
Where the slip velocity is defined as; the contribution of the velocity difference between the line speed and the
particle velocity to explain for the head losses. Mathematically the Stratification Ratio Solids and the Slip
Relative Squared are the same, but physically the Slip Relative Squared tells more about the physics of the
heterogeneous hydraulic transport. So the Srs value explains for the kinetic energy losses in the heterogeneous flow
regime. The potential energy losses are taken into account by the Settling Velocity Hindered Relative, the Shr
value. These potential energy losses are present both in the heterogeneous flow regime and the homogeneous flow
regime.

Many graphs in this book and specifically this chapter have the relative excess hydraulic gradient as the ordinate
and the hydraulic gradient of pure liquid as the abscissa. Since the relative excess hydraulic gradient equals the
mixture hydraulic gradient minus the pure liquid hydraulic gradient, divided by the relative submerged density of
the solids and the volumetric concentration, the graph is almost dimensionless. Almost, because there is are still
some non-linear effects of the relative submerged density and the volumetric concentration. The mixture hydraulic
gradient minus the pure liquid hydraulic gradient is often called the solids effect, so the increase of the hydraulic
gradient due to the presence of solids. The volumetric concentration can be either the spatial or the transport
concentration, depending on the measurement method. Most researchers, in their models, assume the mixture
hydraulic gradient equals the pure liquid hydraulic gradient plus the solids effect. Only the more physical models,
the 2LM and 3 LM models, have a different approach.

Copyright © Dr.ir. S.A. Miedema TOC Page 125 of 304


Introduction Dredging Engineering.

Liquid il curve
Hydraulic gradient im, il vs. Line speed vls
0.40 Fixed Bed Cvs=c.

Sliding Bed Cvs=c.


0.35 Lower Limit
Sliding Bed Cvs=c.
Mean

im-il
0.30
Hydraulic gradient im, il (m water/m)

Sliding Bed Cvs=c.


Upper Limit
Heterogeneous Flow
0.25 Cvs=c.
Equivalent Liquid

im-il
Model
0.20
Homogeneous Flow
Cvs=Cvt=c.

im-il
im-il
Resulting im curve
0.15 Cvs=c.
im-il

Resulting im curve
Cvt=c.
0.10
Limit Deposit Velocity
Cvs=c.
Limit Deposit Velocity
0.05
Cvt=c.
Limit Deposit Velocity

0.00
0 1 2 3 4 5 6 7 8 9 10
Line speed vls (m/sec)
© S.A.M. Dp=0.1524 m, d=1.500 mm, Rsd=1.585, Cv=0.300, μ=0.420

Figure 5-5: The hydraulic gradient im, il and excess hydraulic gradient im-il.
Constant spatial volumetric concentration.

Liquid il curve
Hydraulic gradient im, il vs. Line speed vls
0.40 Fixed Bed Cvs=c.

Sliding Bed Cvs=c.


0.35 Lower Limit
Sliding Bed Cvs=c.
Mean
im-il

0.30
Hydraulic gradient im, il (m water/m)

Sliding Bed Cvs=c.


Upper Limit
Heterogeneous Flow
0.25 Cvs=c.
Equivalent Liquid
im-il

Model
0.20
Homogeneous Flow
Cvs=Cvt=c.
im-il
im-il

im-il

Resulting im curve
0.15 Cvs=c.
Resulting im curve
Cvt=c.
0.10
Limit Deposit Velocity
Cvs=c.
Limit Deposit Velocity
0.05
Cvt=c.
Limit Deposit Velocity

0.00
0 1 2 3 4 5 6 7 8 9 10
Line speed vls (m/sec)
© S.A.M. Dp=0.1524 m, d=1.500 mm, Rsd=1.585, Cv=0.300, μ=0.420
Figure 5-6: The hydraulic gradient im, il and excess hydraulic gradient im-il.
Constant delivered (transport) volumetric concentration.

Figure 5-5 and Figure 5-6 show the solids effect for the constant spatial volumetric concentration Cvs case and the
constant transport volumetric concentration Cvt case. The solids effect in general decreases with increasing line
speed. For low line speeds the Cvt case gives a higher solids effect compared with the Cvs case due to increasing
slip with decreasing line speed.

Page 126 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Delft Head Loss & Limit Deposit Velocity Framework.

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il


Sliding Bed Cvs=c.

1.000
Equivalent Liquid
Relative excess hydraulic gradient Erhg (-)

Model

Homogeneous Flow
Cvs=Cvt=c.

Uniform Sand Cvs=c.


0.100

Uniform Sand Cvt=c.

Limit Deposit Velocity

0.010 Graded Sand Cvs=c.

Graded Sand Cvt=c.

Cv=0.110

0.001
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)
© S.A.M. Dp=0.2032 m, d=0.681 mm, Rsd=1.585, Cv=0.110, μsf=0.416

Figure 5-7: Behavior of narrow graded crushed granite slurry after Clift et al. (1982).

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il


Sliding Bed Cvs=c.

1.000
Equivalent Liquid
Relative excess hydraulic gradient Erhg (-)

Model

Homogeneous Flow
Cvs=Cvt=c.

Uniform Sand Cvs=c.


0.100

Uniform Sand Cvt=c.

Limit Deposit Velocity

0.010 Graded Sand Cvs=c.

Graded Sand Cvt=c.

Cv=0.07-0.16

0.001
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)
© S.A.M. Dp=0.4400 m, d=0.420 mm, Rsd=1.585, Cv=0.110, μsf=0.416

Figure 5-8: Behavior of narrow graded crushed granite slurry after Clift et al. (1982).

Figure 5-7 shows a case where the transition velocity is the same for the collapse of the kinetic interactions and
the start of the lift work. Figure 5-8 shows a case where the transition velocity of the lift work is higher than the
transition velocity for the collapse of the kinetic interactions. The latter results in a collapse of the relative excess
hydraulic gradient. In both examples the same solids are used, but in the latter case the pipe diameter is bigger.
Other experiments by Clift et al. (1982) with narrow graded 0.42 mm masonry sand, shows exactly the same
phenomena.

Copyright © Dr.ir. S.A. Miedema TOC Page 127 of 304


Introduction Dredging Engineering.

5.1.3. Starting Points.


Before discussing the Delft Head Loss & Limit Deposit Velocity (DHLLDV) Framework in detail, some starting
points have to be pointed out. First of all, the Framework is based on a set of 5 sub-models for 5 main flow regimes.
These sub-models are all based on a constant spatial volumetric concentration Cvs. Curves for constant volumetric
transport concentration Cvt are derived from the 5 sub-models based on the slip velocity vsl. The slip velocity vsl is
defined as the difference between the velocity of the mixture vls and the velocity of the solids vs:

 vs   C vt 
v s l  v ls  v s  v ls   1    v ls   1   (5-1)
 v ls   C vs 

For a certain control volume the volumetric transport concentration Cvt can be determined if the volumetric spatial
concentration Cvs and the slip velocity vsl are known, given a certain line speed vls.

 v sl 
C vt   1    C vs (5-2)
 v ls 

Likewise, for a certain control volume, the volumetric spatial concentration Cvs can be determined if the volumetric
transport concentration Cvt and the slip velocity vsl are known, given a certain line speed vls.

 v ls 
C vs     C vt (5-3)
v
 ls  v sl 

These equations will be used a lot in the following derivations and are considered to be well known. The 5 main
flow regimes are:
1 A Fixed Bed (FB) regime or restricted pipe regime. The behavior of this main flow regime is, the solids form
a bed at the bottom of the pipe. This bed is stationary (fixed), so the liquid has to flow through a restricted
area above the bed, resulting in higher pressure losses. At higher line speeds it is probable that part of the
solids start eroding and be transported heterogeneously above the bed. At the Limit Deposit Velocity, the bed
has been eroded completely. As long as the pressure losses correspond with the behavior of flow through the
restricted area above the bed, the flow regime is considered to be a fixed bed regime.
2 A Sliding Bed (SB) regime or sliding friction dominated regime. The behavior of this main flow regime is,
the solids form a sliding bed at the bottom of the pipe. The pressure losses are the sum of the losses as a result
of the sliding friction of the solids and the viscous friction of the liquid. At higher line speeds it is probable
that part of the solids start eroding and be transported heterogeneously above the bed. At the Limit Deposit
Velocity, the bed has been eroded completely. At higher concentration it is possible that sheet flow occurs
and the sliding bed curve is followed right of the intersection with the heterogeneous transport curve. As long
as the pressure losses correspond with the behavior of sliding friction, the pressure loss curves are parallel
with the clean water resistance curve in the im versus vls plot, the sliding bed regime is considered.
3 Heterogeneous (He) transport or collision dominated regime. The behavior of this main flow regime is, the
solids interact with the pipe wall through collisions. The solids are distributed non-uniformly over the cross
section of the pipe with higher concentrations at the bottom of the pipe. This may be due to saltation or to
Brownian motions of the particles in turbulent transport. For very small particles this may follow the fixed
bed regime directly, for coarse particles this will follow the sliding bed regime.
4 Homogeneous (Ho) transport. The behavior of this main flow regime is, the particles are uniformly distributed
over the cross section of the pipe due to the mixing capability of the turbulent flow. The pressure losses behave
according to Darcy Weisbach, but with the mixture density as the liquid density. For very fine particles the
viscosity has to be adjusted by the apparent viscosity.
5 The Sliding Flow (SF) regime. If the ratio between the particle diameter and the pipe diameter is above a
certain value and the spatial volumetric concentration is above about 5%, the turbulence is not capable of
carrying the particles anymore. This will result in a high speed flow with the characteristics of sliding friction,
however the bed concentration decreases with increasing line speed. So it’s named Sliding Flow.

Page 128 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Delft Head Loss & Limit Deposit Velocity Framework.

The hydraulic gradient iw (for water) or il (for a liquid in general) and for a mixture are:

2 2
pl  l  v ls pm  m  v ls
il  iw   and im   (5-4)
l  g  L 2 g Dp l  g  L 2 g Dp

The Relative Excess Hydraulic Gradient Erhg is the difference between the mixture gradient im (in meters of
carrier liquid column) and the hydraulic gradient il divided by the relative submerged density Rsd and the
volumetric concentration Cvs. This Erhg will also be referred to as the solids effect. The Slip Relative Squared Srs
is the Slip Velocity of a particle vsl divided by the Terminal Settling Velocity of a particle vt squared and this Srs
value is a good indication of the excess pressure losses due to the solids in the heterogeneous regime. The Settling
Velocity Hindered Relative Shr is the ratio between the hindered settling velocity vt·(1-Cvs/κC)β and the line speed
vls, divided by the relative submerged density Rsd and the volumetric concentration Cv. For all regimes the Erhg
value is:

im  il
E rh g  (5-5)
R sd  C vs

In the heterogeneous regime the relation between these parameters is:

im  il
E rh g   S h r  S rs (5-6)
R sd  C vs

Figure 5-9, Figure 5-10, Figure 5-11 and Figure 5-12 show the 5 main flow regimes for small, medium and large
particles in an 0.1524 m (6 inch) pipeline. The abscissa, the horizontal axis, is the line speed vls. The ordinate, the
vertical axis, is the hydraulic gradient of the mixture im. The red solid line is the constant volumetric spatial
concentration Cvs line. The green dashed line the constant volumetric transport concentration Cvt line. The light
brown dashed lines show the sliding bed curves, where the thick line is based on the sliding friction coefficient
and the thin lines give a margin of +/- 12.5% of the sliding friction coefficient. The solid blue line is the pure liquid
hydraulic gradient, the dashed blue line the ELM (Equivalent Liquid Model) curve and the dark brown dashed line
the theoretical homogeneous regime curve. The dotted lines give the Limit Deposit Velocity curves for spatial and
transport concentration.

For very fine particles, the fixed bed regime transits directly to the heterogeneous regime, without the occurrence
of the sliding bed regime. This can be seen in Figure 5-9 because the intersection point is below the sliding bed
curve. The Limit Deposit Velocity is at the transition between the heterogeneous regime and the homogeneous
regime. Although there is some slip above the Limit Deposit Velocity, the slip and thus the difference between the
constant volumetric spatial concentration Cvs lines and the constant volumetric transport concentration Cvt lines
increases with a decreasing line speed at line speeds below the Limit Deposit Velocity. The intersection point
between the fixed bed regime and the heterogeneous regime will be at an increasing Erhg value with an increasing
particle diameter.

For medium particles, Figure 5-10, the intersection point between the fixed bed regime and the heterogeneous
regime lies above the sliding bed regime curve, meaning that the fixed bed regime is followed by the sliding bed
regime, followed by the heterogeneous regime, with increasing line speed. The Limit Deposit Velocity is now
somewhere between the intersection of the sliding bed regime with heterogeneous regime and the heterogeneous
regime with the homogeneous regime. The larger the particle the closer is the Limit Deposit Velocity to the
intersection of the sliding bed regime with heterogeneous regime.

The examples given here are for an 0.1524 m pipe. For other pipe diameters, the sliding bed (constant sliding
friction coefficient) and the homogeneous regime curves, will stay at the same position and do not depend on the
pipe diameter. The fixed bed curve will move to the right with increasing pipe diameter, while the heterogeneous
regime curve will move to the left with increasing pipe diameter. One could also say that both curves move
downwards with an increasing pipe diameter.

The transitions between the main flow regimes are not instantaneous, but gradually. Special attention will be given
to the transition between the heterogeneous regime and the homogeneous regime.

Copyright © Dr.ir. S.A. Miedema TOC Page 129 of 304


Introduction Dredging Engineering.

Liquid il curve
Hydraulic gradient im, il vs. Line speed vls
0.250 Fixed Bed Cvs=c.

0.225 Sliding Bed Cvs=c.


Lower Limit
Sliding Bed Cvs=c.
0.200 Mean
Hydraulic gradient im, il (m water/m)

Sliding Bed Cvs=c.


0.175 Upper Limit
Heterogeneous Flow

Heterogeneous
Cvs=c.
0.150
Equivalent Liquid
Fixed Bed Model
0.125 Homogeneous Homogeneous Flow
Cvs=Cvt=c.
0.100 Resulting im curve
Cvs=c.

0.075 Resulting im curve


Cvt=c.
Limit Deposit Velocity
0.050 Cvs=c.
Limit Deposit Velocity
0.025 Cvt=c.
Limit Deposit Velocity

0.000
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
Line speed vls (m/sec)
© S.A.M. Dp=0.1524 m, d=0.200 mm, Rsd=1.585, Cv=0.300, μ=0.416
Figure 5-9: The 3 main flow regimes for fine particles.

Liquid il curve
Hydraulic gradient im, il vs. Line speed vls
0.50 Fixed Bed Cvs=c.

0.45 Sliding Bed Cvs=c.


Lower Limit
Sliding Bed Cvs=c.
0.40 Mean
Hydraulic gradient im, il (m water/m)

Sliding Bed Cvs=c.


0.35 Homogeneous Upper Limit
Sliding Bed

Heterogeneous

Heterogeneous Flow
Cvs=c.
0.30
Equivalent Liquid
Model
0.25
Homogeneous Flow
Cvs=Cvt=c.
0.20 Resulting im curve
Cvs=c.

0.15 Resulting im curve


Cvt=c.
Limit Deposit Velocity
0.10
Fixed Bed Cvs=c.
Limit Deposit Velocity
0.05 Cvt=c.
Limit Deposit Velocity

0.00
0 1 2 3 4 5 6 7 8 9 10
Line speed vls (m/sec)
© S.A.M. Dp=0.1524 m, d=0.500 mm, Rsd=1.585, Cv=0.300, μ=0.416
Figure 5-10: The 4 main flow regimes for medium particles.

For large particles, Figure 5-11, the behavior is similar to the medium particles, except for the fact that the Limit
Deposit Velocity is at the sliding bed regime, below the intersection point between the sliding bed regime and the
heterogeneous regime. This is possible because in reality this transition is not sharp but gradual.

Page 130 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Delft Head Loss & Limit Deposit Velocity Framework.

Very coarse particles, Figure 5-12, show sliding flow behavior. Turbulence is not capable anymore to bring the
particles in suspension. The behavior is a mix of sliding bed and heterogeneous flow.

Liquid il curve
Hydraulic gradient im, il vs. Line speed vls
0.50 Fixed Bed Cvs=c.

0.45 Sliding Bed Cvs=c.


Lower Limit

Heterogeneous
0.40 Homogeneous Sliding Bed Cvs=c.
Mean
Hydraulic gradient im, il (m water/m)

Sliding Bed Cvs=c.


Sliding Bed
0.35 Upper Limit
Heterogeneous Flow
Cvs=c.
0.30
Equivalent Liquid
Model
0.25
Homogeneous Flow
Cvs=Cvt=c.
0.20 Resulting im curve
Cvs=c.

0.15 Resulting im curve


Cvt=c.

0.10 Fixed Bed Limit Deposit Velocity


Cvs=c.
Limit Deposit Velocity
0.05 Cvt=c.
Limit Deposit Velocity

0.00
0 1 2 3 4 5 6 7 8 9 10
Line speed vls (m/sec)
© S.A.M. Dp=0.1524 m, d=1.500 mm, Rsd=1.585, Cv=0.300, μ=0.420
Figure 5-11: The 4 main flow regimes for coarse particles.

Liquid il curve
Hydraulic gradient im, il vs. Line speed vls
0.50 Fixed Bed Cvs=c.
Sliding Flow

0.45 Sliding Bed Cvs=c.


Lower Limit
Sliding Bed Cvs=c.
0.40 Mean
Hydraulic gradient im, il (m water/m)

Sliding Bed Cvs=c.


Sliding Bed

0.35 Upper Limit


Heterogeneous Flow
Cvs=c.
0.30
Equivalent Liquid
Model
0.25
Homogeneous Flow
Cvs=Cvt=c.
0.20 Resulting im curve
Cvs=c.

0.15 Resulting im curve


Cvt=c.
Limit Deposit Velocity
0.10 Fixed Bed Cvs=c.
Limit Deposit Velocity
0.05 Cvt=c.
Limit Deposit Velocity

0.00
0 1 2 3 4 5 6 7 8 9 10
Line speed vls (m/sec)
© S.A.M. Dp=0.1524 m, d=3.000 mm, Rsd=1.585, Cv=0.300, μ=0.416
Figure 5-12: The 3 main flow regimes for very coarse particles, including sliding flow.

Copyright © Dr.ir. S.A. Miedema TOC Page 131 of 304


Introduction Dredging Engineering.

5.1.4. Approach.
Chapter 7 describes the new Delft Head Loss & Limit Deposit Velocity (DHLLDV) Framework. The DHLLDV
Framework is based on uniform sands or gravels and constant spatial volumetric concentration.

1. An explanation of the Delft Head Loss & Limit Deposit Velocity Framework.
2. A detailed description of the 8 different flow regimes and 6 scenarios is given. The occurrence of flow regimes
depends on the particle to pipe diameter ratio and on the spatial volumetric concentration. Figure 5-13 gives
an example of the different flow regimes occurring depending on the particle diameter and the line speed.
Each pipe diameter and each spatial volumetric concentration requires such a graph.
3. The stationary bed regime without sheet flow and with sheet flow. The stationary bed without sheet flow is
based on a 2 layer model for low line speeds and a 3 layer model for higher line speeds. Usually the bed starts
sliding when there is sheet flow, however for small particles it is possible that there is a direct transition from
the stationary bed regime to the heterogeneous flow regime.
4. The sliding bed regime. The sliding bed is based on a 3 layer model showing an almost constant relative excess
hydraulic gradient equal to the sliding friction coefficient. The sliding bed regime does not always occur. The
larger the particles and the larger the volumetric concentration, the higher the probability of the occurrence of
a sliding bed.
5. The heterogeneous regime. The heterogeneous model is based on energy considerations, resulting in a two
component model, potential energy losses and kinetic energy losses.
6. The homogeneous regime. The homogeneous model is based on the equivalent liquid model (ELM) with a
correction based on a particle free viscous sub layer.
7. The sliding flow regime. The sliding flow model assumes a high speed flow with the macroscopic behavior
of sliding friction and heterogeneous flow. The porosity of the bed increases with the line speed and particles
do not necessarily rest on each other.
8. A new model for the Limit Deposit Velocity is derived, consisting of 5 particle size regions and a lower limit.
This model is based on the ratio of the potential energy of the particles to the total energy in the liquid flow
for small particles and on a limiting small bed for large particles.
9. Based on the LDV a method is shown to construct slip velocity or slip ratio curves from zero line speed to the
LDV. Based on the slip ratio, the constant delivered volumetric concentration curves can be constructed. The
resulting model is compared with models from literature.
10. The concentration distribution. Also based on the LDV, in this case the assumption that at the LDV the
concentration at the bottom of the pipe equals the bed concentration, a new diffusivity approach is developed.
The resulting concentration distributions are compared with experiments.
11. The transition heterogeneous versus homogeneous in detail. The transition from the heterogeneous regime to
the homogeneous regime requires special attention. First of all, the transition line speed gives a good indication
of the operational line speed and allows to compare the DHLLDV Framework with many models from
literature. Secondly, at this transition collisions disappear due to near wall lift, while homogeneous transport
is mobilized due to turbulence.
12. Knowing the slip ratio, the bed height for line speeds below the LDV can be determined. Since the LDV is
defined as the line speed above which a sliding or stationary bed does not exist, below this line speed a bed
does exist. New equations are derived for this.
13. Finally the grading of the Particle Size Distribution (PSD) is discussed. A method is given to construct
resulting head loss, slip velocity and bed height curves for graded sands and gravels.
14. Inclined pipes. In real life often inclined pipes are used. Whether its in the ladder of a CSD, the suction pipe
of a TSHD or an upwards or downwards slope, the hydraulic gradient will differ from a horizontal pipe. The
effect of an inclined pipe is derived both for the hydraulic gradient and for the LDV.

Page 132 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Delft Head Loss & Limit Deposit Velocity Framework.

DHLLDV Flow Regime Diagram


Particle diameter (mm)
0.01 0.1 1 10 100
3.0

Undefined
6.0

2.5

He-Ho 5.0
Sliding Flow (SF)
Homogeneous (Ho)
2.0
Durand Froude number FL (-)

4.0
Heterogeneous SB-SF

Line speed (m/s)


(He)
LDV
1.5
Viscous
Effects
3.0

SB-He
Sliding Bed (SB)
1.0
2.0

FB-SB (LSDV)
FB-He
0.5
1.0
Fixed/Stationary Bed (FB)

0.0 0.0
1.E-05 1.E-04 1.E-03 1.E-02 1.E-01
Particle diameter d (m)
© S.A.M.
Dp=0.1524 m, Rsd=1.585, Cvs=0.175, μsf=0.416
Figure 5-13: An example of a flow regime diagram.

Copyright © Dr.ir. S.A. Miedema TOC Page 133 of 304


Introduction Dredging Engineering.

5.1.5. Nomenclature Introduction.


Cvs Spatial volumetric concentration -
Cvt Delivered (transport) volumetric concentration -
d Particle diameter m
d50 50% passing particle diameter m
Dp Pipe diameter m
Erhg Relative excess hydraulic gradient -
ELM Equivalent Liquid Model -
g Gravitational constant 9.1 m/s2 m/s2
il Liquid hydraulic gradient m/m
iw Water hydraulic gradient m/m
im Mixture hydraulic gradient m/m
ΔL Length of pipe m
LDV Limit Deposit Velocity m/s
Δpl Pressure difference liquid kPa
Δpm Pressure difference mixture kPa
PSD Particle Size Diagram -
R The Wilson stratification ratio -
Rsd Relative submerged density -
Shr Settling velocity Hindered Relative -
Srs Slip velocity Relative Squared -
vls Line speed m/s
vs Velocity solids m/s
vsl Slip velocity m/s
vt Terminal settling velocity m/s
ρl Liquid density ton/m3
ρm Mixture density ton/m3
κC Concentration eccentricity hindered settling -
λl Darcy Weisbach friction factor -
μsf Sliding friction coefficient -
νl Kinematic viscosity m2/s

Page 134 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Delft Head Loss & Limit Deposit Velocity Framework.

5.2. Flow Regimes and Scenario’s.


5.2.1. Introduction.
In dredging, the hydraulic transport of solids is one of the most important processes. Since the 50’s many
researchers have tried to create a physical mathematical model in order to predict the head losses in slurry transport.
We can think of the models of Durand, Condolios, Gibert, Worster, Zandi & Govatos, Jufin Lopatin, Fuhrboter,
Newitt, Doron, Wilson, Matousek, Turian & Yuan and the SRC model. Some models are based on
phenomenological relations and thus result in semi empirical relations, others tried to create models based on
physics, like the two and three layer models. It is however the question whether slurry transport can be modeled
this way at all. Observations in our laboratory show a process which is often non-stationary with respect to time
and space. Different physics occur depending on the line speed, particle diameter, concentration and pipe diameter.
These physics are often named flow regimes; fixed bed with and without sheet flow or suspension, sliding bed,
heterogeneous transport, (pseudo) homogeneous transport and sliding flow. It is also possible that more regimes
occur at the same time, like, a fixed bed in the bottom layer with heterogeneous transport in the top layer.

It is the observation of the authors that researchers often focus on a detail and sub-optimize their model, which
results in a model that can only be applied for the parameters used for their experiments. At high line speeds the
volumetric spatial concentration (volume based) and the volumetric transport concentration (volume flux based)
are almost equal, because all the particles are in suspension with a small slip related to the carrier liquid velocity.
The difference of the head loss between the two concentrations will be within the margin of the scatter of the
experiments. At low line speeds however, there may be a sliding or fixed bed, resulting in a big difference between
the two concentrations and thus between laboratory and real life situations.

This chapter describes 8 flow regimes and 6 possible scenarios.

The flow regimes for constant spatial volumetric concentration Cvs are, from line speed zero with increasing line
speed:
1: Fixed bed without suspension (fine particles) or sheet flow (coarse particles).
2: Fixed bed with suspension (fine particles) or sheet flow (coarse particles).
3: Fixed bed with suspension (fine particles) or sliding bed with sheet flow (coarse particles).
For fine to coarse particles d/Dp<0.015:
5: Heterogeneous transport Cvs≈Cvt.
5/6: Pseudo homogeneous transport, Cvs≈Cvt.
6: Homogeneous transport, Cvs≈Cvt.
For very coarse particles d/Dp>0.015:
7: Sliding flow.

The flow regimes for constant delivered volumetric concentration Cvt are, from line speed zero with increasing
line speed:
8: Fixed bed with suspension (fine particles) or sheet flow (coarse particles).
4: Fixed bed with suspension (fine particles) or sliding bed with sheet flow (coarse particles).
For fine to coarse particles d/Dp<0.015:
5: Heterogeneous transport Cvs≈Cvt.
5/6: Pseudo homogeneous transport, Cvs≈Cvt.
6: Homogeneous transport, Cvs≈Cvt.
For very coarse particles d/Dp>0.015:
7: Sliding flow.

3 scenarios are based on a constant volumetric spatial concentration (usually in a laboratory) and 3 scenarios are
based on a constant volumetric transport concentration (usually in real life). The flow regimes and scenarios are
explained and examples of experiments are given. Based on the experimental evidence, one can conclude that the
approach followed in this book gives a good resemblance with the reality.

Copyright © Dr.ir. S.A. Miedema TOC Page 135 of 304


Introduction Dredging Engineering.

5.2.2. Concentration Considerations.


Based on an analysis of many experiments from literature, 8 flow regimes and 6 scenarios can be distinguished,
which will be discussed in the next chapters. In order to understand these 8 flow regimes and 6 scenarios, the
difference between the spatial volumetric concentration Cvs and the volumetric transport (delivered) concentration
Cvt will first be discussed. In hydraulic transport, 2 definitions of the concentration are often used. Contractors are
interested in the delivered volumetric concentration, also named the volumetric transport concentration Cvt. Cvt is
defined as the ratio between the volume flow of solids and the volume flow of the mixture. In general one can say
that the average solids velocity will be smaller than the average mixture velocity. The difference is called the slip
velocity. The spatial volumetric concentration Cvs is defined as the volume of solids divided by the volume of the
mixture containing these solids. So the spatial volumetric concentration is based on a volume ratio, while the
delivered volumetric concentration is based on a volume flux ratio. Concentration (Cvs) is usually derived from
density meter or U-loop readings as:

m  l m  l
C vs  ( d e n s it y m e t e r ) or C vt  ( U - lo o p ) (5-7)
s  l s  l

A radioactive density meter reads a density of the entire mass of slurry in the pipe, and thus is best suited to
measuring Cvs. In addition the placement of the meter in horizontal or vertical pipe, can affect the readings. A U-
tube device reads the delivered density and is thus best suited to measuring Cvt. In a closed loop system, we will
know the volume of the closed loop and amount of material added, and thus can calculate Cvs directly, but not
necessarily Cvt. The volumetric delivered (transport) concentration is:

Vs Qs vs  A s v s  C vs  A p vs
C vt      C vs  (5-8)
Vm Qm vm  A p vm  A p vm

With vs the average velocity of the solids and vm the average velocity of the mixture, also called the line speed vls.
The volumetric spatial concentration is based on the volume ratio solids/mixture according to:

Vs
C vs  (5-9)
Vm

The slip velocity vsl is defined as the difference between the velocity of the mixture vm and the velocity of the
solids vs:

 vs   C vt 
v sl  v m  v s  v m   1    vm  1   (5-10)
 vm   C vs 

Because of the fact that most experiments are carried out in a closed loop system, the concentration might be
determined by the ratio of the volume of solids divided by the volume of the closed loop system.

Vs
C vs  (5-11)
Vcl

This means that the concentration of solids in the liquid above the bed will be much smaller once a bed is formed.
Now assume a bed with a porosity n of about 40% containing 50% of the solids, matching the v50 of Wilson (1997).
This gives for the total bed volume in the closed loop system:

Vs 1 C vs  Vcl
Vb    (5-12)
2 1 n 2  1  n 

The volume of solids in suspension is the same, so the volume of the solids in the liquid is:

Vs
V s ,s  (5-13)
2

Page 136 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Delft Head Loss & Limit Deposit Velocity Framework.

The volume of the mixture in suspension above the bed equals the closed loop volume minus the bed volume.

C vs  Vcl  2   1  n   C vs 
V m ,s  V c l  V b  V c l   Vcl    (5-14)
2  1  n   2  1  n  
 

The concentration of the solids in suspension is the volume of these solids, divided by the volume of the closed
loop system minus the volume of the bed.

Vs C vs  1  n   C vs 0 .6  C v s C vs
C v s ,s     
2  V m ,s  2   1  n   C vs  2   1  n   C vs 1 .2  C v s 2  1 .6 6  C v s (5-15)
2 
 2  1  n 
 

Of course the closed loop will not consist of just horizontal parts where a bed may occur, but the above example
is just meant to give an indication.

This implies that at low spatial volumetric concentrations Cvs, the concentration in the suspension phase, the
heterogeneous transport phase, is 50% of the total volumetric concentration. At a high concentration of Cvs=0.3,
the concentration of the heterogeneous phase is still reduced to 0.2. At a concentration of Cvs=0.6, the above
equation results in a concentration of 0.6, which makes sense, since this is solid sand and there is no suspension
anymore. When experiments are carried out it should be clear which concentration is used. Is it the concentration
based on the volume of the closed loop system, giving some constant volumetric spatial concentration? Is it the
concentration based on radio active density meters in the pipe section where also the hydraulic gradient is
measured, resulting in a spatial volumetric concentration? Or is the concentration measured with a U-tube resulting
in a volumetric transport concentration.

Now in a real life production situation there is not a closed loop system, but an open system. There is not a fixed
amount of solids in the pipeline, which can be divided in a part in a bed and the rest in suspension. Instead, the
supply at the suction mouth can vary from water to twice or more that the delivered concentration. In a stable
situation, the production that enters the system is equal to the production that leaves the system. The concentration
is determined at the suction mouth and although there may be a bed in part of the pipeline, this does not change
the transport concentration, it just increases the line speed and concentration above the bed compared with a
pipeline without a bed, due to the conservation of volume in the pipeline. The conclusion of the above
considerations is that for a good interpretation of the results of experiments, the method of determining the
concentration should be known. It is also important how the results are presented. Graf & Robinson (1970) for
example, present their results based on a constant amount of solids in their closed loop system, while Doron &
Barnea (1987) connect data points with constant volumetric transport concentration. The presentations of the
results are different, while the physics are the same.

Figure 5-14: The definitions for fully stratified flow.

Copyright © Dr.ir. S.A. Miedema TOC Page 137 of 304


Introduction Dredging Engineering.

5.2.3. The 8 Flow Regimes Identified.


In literature different flow patterns or flow regimes are distinguished. Durand & Condolios (1952) distinguish 4
regimes, based on the particle size. Abulnaga (2002) also distinguished 4 regimes based on the actual flow of
particles and their size. Matousek (2004) in his lecture notes distinguishes 6 flow regimes. Here we will consider
8 flow regimes and 6 scenarios for laboratory and real life conditions. These are (Figure 5-14 gives some
definitions of fully stratified flow in a pipe):

Table 5.2-1: The 8 possible flow regimes.

1: Fixed bed without suspension or sheet flow, constant Cvs.

Under laboratory circumstances with a constant spatial volumetric concentration Cvs, at low line speeds all the
particles are in a stationary (fixed) bed at the bottom of the pipe. Above the bed the liquid is flowing through a
smaller cross-section A1=Ap-A2. This gives a higher effective line speed vls,e=vls·Ap/A1. Since the bottom of
this cross–section consists of particles, the resulting Darcy-Weisbach friction factor λm has to be determined by
taking a weighted average of the friction factor of the liquid-bed interface λ12 and the friction factor of the
liquid-pipe wall interface λl. The method of Miedema & Matousek (2014) can be applied to determine the
Darcy-Weisbach friction coefficient λ12 on the liquid-bed interface. This method does distinguish between fine
and coarse particles. The particle diameter influences the Darcy-Weisbach friction coefficient λ12 by using the
particle diameter as the bed roughness.

The Shields parameter is below a critical Shields value of about 1 so no sheet flow or suspension occurs.

The total pressure loss is thus determined by viscous friction on the liquid-bed interface and the liquid-
pipe wall interface, the spatial volumetric concentration and the particle diameter.

2: Fixed bed with suspension or sheet flow, constant Cvs.

Under laboratory circumstances with a constant spatial volumetric concentration Cvs, at medium-low line speeds
most of the particles are in a stationary (fixed) bed at the bottom of the pipe. Above the bed a suspension (fine
particles) or a sheet flow (coarse particles) is flowing through a smaller cross-section A1=Ap-A2. This gives a
higher effective line speed vls,e=vls·Ap/A1. Since the bottom of this cross–section consists of particles, the Darcy-
Weisbach friction factor λm has to be determined by taking a weighted average of the friction factor of the
liquid-bed interface λ12 and the friction factor of the liquid-pipe wall interface λl. The method of Miedema &
Matousek (2014) for sheet flow can be applied to determine the friction coefficient λ12 on the liquid-bed
interface. This method does distinguish between fine and coarse particles and gives an explicit relation for the
Darcy-Weisbach friction factor. The particle diameter influences the Darcy-Weisbach friction coefficient λ12
slightly.

Page 138 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Delft Head Loss & Limit Deposit Velocity Framework.

The Shields parameter is above a critical Shields value of about 1, so sediment transport/erosion occurs
in the form of sheet flow or suspension.

The total pressure loss is thus determined by viscous friction, shear stresses in the sheet flow layer, the
spatial volumetric concentration and the particle diameter.

3: Fixed bed with suspension or sliding bed with sheet flow, constant Cvs.

Under laboratory circumstances with a constant spatial volumetric concentration Cvs, for coarse particles the
bed is sliding with sheet flow at the top, where the thickness of the sheet flow layer increases with an increasing
velocity difference between the flow above the bed and the bed, while for fine particles the shear stress on the
bed is not high enough to make it start sliding, but more and more particles will be in suspension as the line
speed increases. For fine particles the behavior starts following the heterogeneous behavior more and more with
increasing line speed.

Since coarse particles in sheet flow require an upwards force at least equal to their submerged weight, which
results from interparticle forces, an equal downwards force will act on the bed-sheet flow layer interface. The
result is a total normal force between the bed and the pipe wall of about the submerged weight of the particles.
From experiments it appears that this normal force is almost a constant times the spatial volumetric
concentration, according to the Newitt et al. (1955) model. This vertical force times the friction coefficient μsf
determines the sliding friction force. The friction coefficient μsf will have a value of about 0.416, but should
preferably be determined by experiments, since it is a property of the particles depending on the shape of the
particles.

The Shields parameter is above a critical Shields value of 1, so sediment transport/erosion occurs.

The total pressure loss is thus determined by sliding friction between the bed and the pipe wall for coarse
particles and by energy losses due to collisions for fine particles.

4: Fixed bed with suspension or sliding bed with sheet flow, constant Cvt.

Under laboratory circumstances with a constant spatial volumetric concentration Cvs, for coarse particles the
bed is sliding with sheet flow, where the thickness of the sheet flow layer increases further with an increasing
velocity difference between the flow above the bed and the bed, while for fine particles the shear stress on the
bed is not high enough to make it start sliding, but almost all particles will be in suspension as the line speed
increases. For fine particles the behavior starts following the heterogeneous behavior as the line speed increases.

Since coarse particles in sheet flow require an upwards force at least equal to their submerged weight, which
results from interparticle forces, an equal downwards force will act on the bed. The result is a total normal force
between the bed and the pipe wall of about the submerged weight of the particles. From experiments it appears

Copyright © Dr.ir. S.A. Miedema TOC Page 139 of 304


Introduction Dredging Engineering.

that this normal force is almost a constant times the spatial volumetric concentration, according to the Newitt
et al. (1955) model. This vertical force times the friction coefficient μsf determines the sliding friction force.
The friction coefficient μsf will have a value of about 0.416, but should preferably be determined by
experiments, since it is a property of the particles depending on the shape of the particles.

The difference with flow regime 3 is, that here the transport/delivered concentration is known. With decreasing
line speed and constant delivered volumetric concentration, the spatial volumetric concentration is increasing.
The spatial concentration will be higher than the delivered concentration, resulting in a higher resistance. So
the constant Cvt curve will always be higher than the constant Cvs curve. The difference increases with
decreasing line speed.

The Shields parameter is above a critical Shields value of 1, so sediment transport/erosion occurs.

The total pressure loss is thus determined by sliding friction between the bed and the pipe wall for coarse
particles and by energy losses due to collisions for fine particles.

5: Heterogeneous transport, Cvt≈Cvs.

When the line speed increases further, the difference between spatial and delivered concentration becomes
smaller. At a certain line speed the macroscopic behavior changes from stationary or sliding bed behavior to
heterogeneous behavior.

The turbulent forces interacting with the particles are not strong enough to create a uniform distribution
throughout the cross-section of the pipe. A definite concentration gradient exists along the vertical profile of
the pipe with the highest concentration at the bottom. There may still be deposits, but most particles move in
suspension or a sort of sheet flow layer. There is however an interaction between the particles and the bottom
of the pipe. These interactions, collisions, cause the loss of kinetic energy of the particles and are the main cause
of the pressure losses. Since the number of collisions per unit of time depends mainly on the terminal settling
velocity of the particles, it will be almost constant, resulting in pressure losses that are reversely proportional
with the line speed or the line speed to a higher power. Since the particles move up and down in the pipe, based
on the terminal settling velocity and hindered settling, there are also potential energy losses. The pressure losses
can be determined according to Durand & Condolios (1952), Jufin & Lopatin (1966), Miedema & Ramsdell
(2013) or others. The heterogeneous model is the same for fine and coarse particles, but the line speed range
where it occurs depends on the particle size.

The Shields parameter is very high above the Shields curve, resulting in suspension/saltation.

The total pressure loss is determined by potential and kinetic energy losses.

Page 140 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Delft Head Loss & Limit Deposit Velocity Framework.

6: Homogeneous transport, Cvt≈Cvs.

When the line speed increases further, the difference between spatial and delivered concentration becomes
smaller and smaller. The turbulent forces interacting with the particles are strong enough to create a uniform
distribution throughout the cross-section of the pipe.

The turbulent forces interacting with the particles are so strong that the mixture has an almost uniform
distribution throughout the cross-section of the pipe. True homogeneous flows is not possible, since for the
turbulent forces to overcome gravity, a concentration gradient has to exist. Pseudo homogeneous regimes
usually occur with very fine particles or at very high line speeds. The pressure losses in this regime can be
modeled using the adapted/modified equivalent liquid model (ELM). It is assumed that the spatial volumetric
concentration Cvs and the volumetric transport (delivered) concentration Cts are almost equal.

The Shields parameter is very high above the Shields curve, resulting in a suspension.

The total pressure loss is determined by the work carried out by lift forces and turbulent dispersion.

5/6: Pseudo homogeneous transport, Cvt≈Cvs.

At the line speed where heterogeneous and homogeneous transport meet, there will be a transition between the
two regimes. If the turbulent near wall lift force equals the submerged weight of the particle, this lift force will
prevent the particles from hitting the bottom of the pipe, resulting in a sudden drop of the heterogeneous pressure
losses. At slightly higher line speeds the lift force is strong enough to push the particles into the turbulent flow,
where turbulent dispersion will take care of further mixing. In between there may be a gap resulting in almost
no additional pressure losses. This occurs for particles with diameter from 0.1-0.5 mm with bigger pipe
diameters. The pressure losses can be determined according to Miedema &Ramsdell (2013).

The Shields parameter is very high above the Shields curve, resulting in a suspension.

The total pressure loss is determined by decreasing potential and kinetic energy losses and by increasing
equivalent liquid model (ELM) behavior.

7: Sliding Flow.

At relatively low concentrations and relatively small particle diameters, the sliding bed regime will have a
transition to the heterogeneous regime at the intersection of the two regimes. This is the result of lift forces
strong enough to lift the particles and turbulent dispersion to mix them into a heterogeneous mixture. However
when the weight of the bed is bigger than the total lift forces, this will not occur and the particles stay in the
bed in a sort of sheet flow. A second reason may be that at high concentrations the space above the bed is not
big enough to fully develop turbulence. The pressure losses in this regime are much higher than the pressure
losses with heterogeneous transport at lower concentrations. The pressure losses can be determined according
to Miedema &Ramsdell (2013). The term Sliding Flow is chosen, because there is flow but the flow resistance
has the character of sliding friction.

The Shields parameter is far above the Shields curve, so sediment transport/erosion occurs.

The total pressure loss is thus determined by a combination of sliding friction between the bed and the
pipe wall and kinetic and potential energy losses. The larger the particle diameter to pipe diameter ratio,
the more this tends to sliding friction behavior and the smaller the heterogeneous contribution.

Copyright © Dr.ir. S.A. Miedema TOC Page 141 of 304


Introduction Dredging Engineering.

8: Fixed bed with suspension, constant Cvt.

Under real life conditions, there will be a “constant” volumetric transport concentration with decreasing line
speed. There will be equilibrium between erosion and deposition, resulting in a certain bed height. Gibert (1960)
has proposed that the Froude number will be equal to the Froude number at the Limit Deposit Velocity. In this
case, the Limit Deposit Velocity is defined as the velocity where the sliding bed has vaporized due to erosion.
With decreasing line speed, the bed height increases and so do the pressure losses. Once the bed height is
known, the pressure losses can be determined according to the Newitt et al. (1955) model. This regime occurs
if the relative excess hydraulic gradient is high enough to result in a sliding bed and so this will occur much
more with small pipe diameters then with large pipe diameter.

The Shields parameter is above the a critical Shields value, so erosion occurs.

The total pressure loss is thus determined by sliding friction between the bed and the pipe wall, where
the spatial concentration is increasing with decreasing line speed, while the transport concentration is a
constant.

5.2.4. The 6 Scenario’s Identified.


In pipes with small diameters the hydraulic gradient will be relatively high, resulting in relatively high hydraulic
gradients when transporting a mixture. This results in hydraulic gradients approaching the hydraulic gradient
required to create a sliding bed. In pipes with large diameters the hydraulic gradient will be relatively small, also
resulting in relatively small hydraulic gradients when transporting a mixture. This results in hydraulic gradients
too small compared with the hydraulic gradient required to create a sliding bed.

From the 8 flow regimes, 6 scenarios can be constructed, where a scenario does not have to contain all 8 flow
regimes. A scenario describes the flow regime behavior when the line speed increases from zero to a certain
maximum. This maximum is arbitrary but should be related to practical line speeds. So this maximum increases
with increasing pipe diameter.

The 8 flow regimes and 6 scenarios are shown in Figure 5-15, Figure 5-16, Figure 5-17, Figure 5-18, Figure 5-19
and Figure 5-20. Figure 5-15 and Figure 5-16 show the scenario’s L1 for laboratory conditions and R1 for real life
conditions for fine sands. Figure 5-17 and Figure 5-18 show the scenario’s L2 for laboratory conditions and R2
for real life conditions for coarse sands. Figure 5-19 and Figure 5-20 show the scenario’s L3 for laboratory
conditions and R3 for real life conditions for gravels.

The difference between laboratory conditions and real life conditions can be found at low line speeds where the
volumetric spatial Cvs and transport Cvt concentrations differ substantially due to slip. At higher line speeds with
heterogeneous and (pseudo) homogeneous transport it is assumed that the slip velocity vsl is small compared to the
line speed vls.

Most of the graphs are made for sands and gravels with water as the carrier liquid, however the modelling is also
valid for other solids and other liquids.

Page 142 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Delft Head Loss & Limit Deposit Velocity Framework.

5.2.4.1. Scenarios L1 & R1.

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il


Fixed Bed Cvs=c.

Sliding Bed Cvs=c.


Lower Limit
Sliding Bed Cvs=c.
Mean
Relative excess hydraulic gradient Erhg (-)

Heterogeneous
1.000 8 Sliding Bed Cvs=c.
Upper Limit
Heterogeneous Flow
Cvs=c.
Equivalent Liquid
Homogeneous Model
Homogeneous Flow
Cvs=Cvt=c.
6 Resulting Erhg curve
2
Cvs=c.
3 Resulting Erhg curve
Cvt=c.
Fixed Bed Limit Deposit Velocity
Cvs=c.
0.100 Limit Deposit Velocity
4
Cvt=c.
Limit Deposit Velocity

Ratio Potential/Kinetic
5
Energy
Fixed Bed Cvs=c.
1
Sliding Bed Cvs=c.
Lower Limit
Sliding Bed Cvs=c.
Mean
Sliding Bed Cvs=c.
0.010 Upper Limit
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)
© S.A.M. Dp=0.1524 m, d=0.200 mm, Rsd=1.585, Cv=0.300, μ=0.416

Figure 5-15: The definition of the pressure losses, scenario’s L1 and R1, E rhg(il).

Table 5.2-2: Scenario’s L1 and R1.


Scenario L1 (the red solid line)

Table 5.2-3: Indication of occurrence of L1.


Dp d Cv
<< >> >>
< > >
- - -
> < <
>> << <<

Starting at a line speed vls=0, there will be a stationary (fixed) bed (1). When the line speed is increased, there
will not be erosion until the Shields parameter is high enough above the Shields curve. Increasing the line speed
further will result in erosion and suspension or sheet flow of the particles (2). The upwards directed solid red
curve becomes steeper.

At a certain line speed, the macroscopic behavior will have a transition from fixed bed to heterogeneous (3).
This also means that the excess pressure losses go from shear stress dominated to collision dominated.
Increasing the line speed further results in heterogeneous transport (5). The solid red curve is downwards
directed.

The Limit Deposit Velocity is somewhere in the heterogeneous flow regime.

Increasing the line speed further, results in a transition region between heterogeneous transport and (pseudo)
homogeneous transport (5/6). At very high line speeds, the regime will be the homogeneous regime (6).
Whether this regime will be reached with practical line speeds depends completely on the combination of the
parameters involved. The solid red curve is upwards directed again.

Copyright © Dr.ir. S.A. Miedema TOC Page 143 of 304


Introduction Dredging Engineering.

Scenario R1 (the green dashed line)

Table 5.2-4: Indication of occurrence of R1.


Dp d Cv
<< >> >>
< > >
- - -
> < <
>> << <<

Starting at a line speed vls=0, there will be equilibrium between erosion and deposition, resulting in a certain
bed height. Above the bed there will be heterogeneous transport or suspension. At very low line speeds, the
hydraulic gradient is so high that a sliding bed may occur (8). At these low line speeds, it is also possible that
only part of the bed is siding resulting in sliding friction that is not fully mobilized. The green dashed curve is
slightly downwards directed. The height of the green dashed line depends on the Cvt. A smaller Cvt gives a
higher green dashed line, because a smaller Cvt will have more slip. The red solid line (constant Cvs) is hardly
influenced by the spatial concentration.

At higher line speeds, the bed has a higher velocity and starts vaporizing, the hydraulic gradient drops, resulting
in a transition towards heterogeneous behavior (4) and at higher line speeds full heterogeneous behavior (5).
The dashed green line is downwards directed.

The Limit Deposit Velocity is somewhere in the heterogeneous flow regime.

Increasing the line speed further, results in a transition region between heterogeneous transport and (pseudo)
homogeneous transport (5/6). At very high line speeds, the regime will be the homogeneous regime (6).
Whether this regime will be reached with practical line speeds depends completely on the combination of the
parameters involved. The green dashed line is upwards directed again. The solid red line and the dashed green
line coincide.

Liquid il curve
Hydraulic gradient im, il vs. Line speed vls
0.250 Fixed Bed Cvs=c.

0.225 Sliding Bed Cvs=c.


8 Lower Limit
Sliding Bed Cvs=c.
0.200 Mean
Hydraulic gradient im, il (m water/m)

Sliding Bed Cvs=c.


0.175 Upper Limit
Heterogeneous Flow
Heterogeneous

Cvs=c.
0.150
Equivalent Liquid
Fixed Bed Model
0.125 Homogeneous Homogeneous Flow
6 Cvs=Cvt=c.
3
0.100 2 Resulting im curve
Cvs=c.

0.075 Resulting im curve


Cvt=c.
4
Limit Deposit Velocity
0.050 5 Cvs=c.
Limit Deposit Velocity
0.025 Cvt=c.
1 Limit Deposit Velocity

0.000
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
Line speed vls (m/sec)
© S.A.M. Dp=0.1524 m, d=0.200 mm, Rsd=1.585, Cv=0.300, μ=0.416

Figure 5-16: The definition of the pressure losses, scenario’s L1 and R1, im(vls).

Page 144 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Delft Head Loss & Limit Deposit Velocity Framework.

5.2.4.2. Scenarios L2 & R2.

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il


Fixed Bed Cvs=c.

Sliding Bed
Sliding Bed Cvs=c.
Lower Limit
Relative excess hydraulic gradient Erhg (-)

Sliding Bed Cvs=c.


1.000 8 Mean
Sliding Bed Cvs=c.
3 4 Upper Limit

Homogeneous Heterogeneous Flow


Cvs=c.
6 Equivalent Liquid
2
Heterogeneous Model
Homogeneous Flow
Fixed Bed 5 Cvs=Cvt=c.
Resulting Erhg curve
0.100 Cvs=c.
Resulting Erhg curve
Cvt=c.
Limit Deposit Velocity
1 Cvs=c.
Limit Deposit Velocity
Cvt=c.
Limit Deposit Velocity

Ratio Potential/Kinetic
Energy
0.010
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)
© S.A.M. Dp=0.1524 m, d=1.500 mm, Rsd=1.585, Cv=0.300, μ=0.420

Figure 5-17: The definition of the pressure losses, scenario’s L2 and R2, Erhg(il).

Table 5.2-5: Scenario’s L2 and R2.


Scenario L2 (the red solid line)

Table 5.2-6: Indication of occurrence of L2.


Dp d Cv
<< >> >>
< > >
- - -
> < <
>> << <<

Starting at a line speed vls=0, there will be a stationary (fixed) bed (1). When the line speed is increased, there
will not be erosion until the Shields parameter is high enough above the Shields curve. Increasing the line speed
further will result in erosion and suspension or sheet flow of the particles (2). The upwards directed solid red
curve becomes steeper.

At a certain line speed, the hydraulic gradient is high enough to make the bed to start sliding (3). Increasing the
line speed further will result in an increase of the velocity of the bed and an increase of the erosion. The relative
excess hydraulic gradient remains constant, because the weight of the suspension and the bed is a constant,
resulting in an almost constant friction force. Increasing the line speed further will give a transition to
heterogeneous transport (5). The solid red line is first horizontal and then downwards directed.

The Limit Deposit Velocity is somewhere at the right of the sliding bed regime.

Increasing the line speed further, results in a transition region between heterogeneous transport and (pseudo)
homogeneous transport (5/6). At very high line speeds, the regime will be the homogeneous regime (6).
Whether this regime will be reached with practical line speeds depends completely on the combination of the
parameters involved. The solid red curve is upwards directed again.

Copyright © Dr.ir. S.A. Miedema TOC Page 145 of 304


Introduction Dredging Engineering.

Scenario R2 (the green dashed line)

Table 5.2-7: Indication of occurrence of R2.


Dp d Cv
<< >> >>
< > >
- - -
> < <
>> << <<

Starting at a line speed vls=0, there will be equilibrium between erosion and deposition, resulting in a certain
bed height. Above the bed there will be heterogeneous transport or suspension. At very low line speeds, the
hydraulic gradient is so high that a sliding bed may occur (8). At these low line speeds, it is also possible that
only part of the bed is siding resulting in sliding friction that is not fully mobilized. The green dashed curve is
slightly downwards directed. The height of the green dashed line depends on the Cvt. A smaller Cvt gives a
higher green dashed line, because a smaller Cvt will have more slip. The red solid line (constant Cvs) is hardly
influenced by the spatial concentration.

At higher line speeds, the bed has a higher velocity and starts vaporizing, the hydraulic gradient drops, resulting
in a transition towards heterogeneous behavior (4) and at higher line speeds full heterogeneous behavior (5).
The dashed green line is downwards directed.

The Limit Deposit Velocity is somewhere at the right of the sliding bed regime.

Increasing the line speed further, results in a transition region between heterogeneous transport and (pseudo)
homogeneous transport (5/6). At very high line speeds, the regime will be the homogeneous regime (6).
Whether this regime will be reached with practical line speeds depends completely on the combination of the
parameters involved. The solid red curve is upwards directed again.

Liquid il curve
Hydraulic gradient im, il vs. Line speed vls
0.50 Fixed Bed Cvs=c.

0.45 Sliding Bed Cvs=c.


Lower Limit
Heterogeneous

0.40 Homogeneous 6 Sliding Bed Cvs=c.


Mean
8
Hydraulic gradient im, il (m water/m)

Sliding Bed Cvs=c.


Sliding Bed

0.35 Upper Limit


Heterogeneous Flow
Cvs=c.
0.30
Equivalent Liquid
Model
0.25
Homogeneous Flow
3 Cvs=Cvt=c.
4
0.20 5 Resulting im curve
Cvs=c.
Resulting im curve
0.15 2 Cvt=c.

0.10 Fixed Bed Limit Deposit Velocity


Cvs=c.
Limit Deposit Velocity
0.05 Cvt=c.
1
Limit Deposit Velocity

0.00
0 1 2 3 4 5 6 7 8 9 10
Line speed vls (m/sec)
© S.A.M. Dp=0.1524 m, d=1.500 mm, Rsd=1.585, Cv=0.300, μ=0.420

Figure 5-18: The definition of the pressure losses, scenario’s L2 and R2, im(vls).

Page 146 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Delft Head Loss & Limit Deposit Velocity Framework.

5.2.4.3. Scenarios L3 & R3.

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il


Fixed Bed Cvs=c.

Sliding Bed

Sliding Flow
Sliding Bed Cvs=c.
Lower Limit
Relative excess hydraulic gradient Erhg (-)

Sliding Bed Cvs=c.


1.000 8 Mean
Sliding Bed Cvs=c.
8 7 5 6 Upper Limit
Heterogeneous Flow
Cvs=c.

Homogeneous
Equivalent Liquid
2 Model
Fixed Bed Homogeneous Flow
Cvs=Cvt=c.
Resulting Erhg curve
0.100 Cvs=c.
Resulting Erhg curve
Cvt=c.
1 Limit Deposit Velocity
Cvs=c.
Limit Deposit Velocity
Cvt=c.
Limit Deposit Velocity

Ratio Potential/Kinetic
Energy
0.010
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)
© S.A.M. Dp=0.1524 m, d=3.000 mm, Rsd=1.585, Cv=0.300, μ=0.416

Figure 5-19: The definition of the pressure losses, scenario’s L3 and R3, Erhg(il).

Table 5.2-8: Scenario’s L3 and R3.


Scenario L3 (the red solid line)

Table 5.2-9: Indication of occurrence of L3.


Dp d Cv
<< >> >>
< > >
- - -
> < <
>> << <<

Starting at a line speed vls=0, there will be a stationary (fixed) bed (1). When the line speed is increased, there
will not be erosion until the Shields parameter is high enough above the Shields curve. Increasing the line speed
further will result in erosion and suspension or sheet flow of the particles (2). The upwards directed solid red
curve becomes steeper.

At a certain line speed, the hydraulic gradient is high enough to make the bed to start sliding (3). Increasing the
line speed further will result in an increase of the velocity of the bed and an increase of the erosion. The relative
excess hydraulic gradient remains constant, because the weight of the suspension and the bed is a constant,
resulting in an almost constant friction force. Increasing the line speed further will give a transition to sliding
flow (7). The solid red line is first horizontal and then slightly downwards directed.

The Limit Deposit Velocity is somewhere at the right of the sliding bed regime or in the sliding flow regime.

Increasing the line speed further, results in a transition region between sliding flow and (pseudo) homogeneous
transport (5/6). At very high line speeds, the regime will be the homogeneous regime (6). Whether this regime
will be reached with practical line speeds depends completely on the combination of the parameters involved.
The solid red curve is upwards directed again.

Copyright © Dr.ir. S.A. Miedema TOC Page 147 of 304


Introduction Dredging Engineering.

Scenario R3 (the green dashed line)

Table 5.2-10: Indication of occurrence of R3.


Dp d Cv
<< >> >>
< > >
- - -
> < <
>> << <<

Starting at a line speed vls=0, there will be equilibrium between erosion and deposition, resulting in a certain
bed height. Above the bed there will be heterogeneous transport or suspension. At very low line speeds, the
hydraulic gradient is so high that a sliding bed may occur (8). At these low line speeds, it is also possible that
only part of the bed is siding resulting in sliding friction that is not fully mobilized. The green dashed curve is
slightly downwards directed. The height of the green dashed line depends on the Cvt. A smaller Cvt gives a
higher green dashed line, because a smaller Cvt will have more slip. The red solid line (constant Cvs) is hardly
influenced by the spatial concentration.

At higher line speeds, the bed has a higher velocity and starts vaporizing, the hydraulic gradient drops, resulting
in a transition towards sliding flow behavior (7) and at higher line speeds full heterogeneous behavior (5). The
dashed green line is slightly downwards directed.

The Limit Deposit Velocity is somewhere at the right of the sliding bed regime or in the sliding flow regime.

Increasing the line speed further, results in a transition region between heterogeneous transport and (pseudo)
homogeneous transport (5/6). At very high line speeds, the regime will be the homogeneous regime (6).
Whether this regime will be reached with practical line speeds depends completely on the combination of the
parameters involved. The solid red curve is upwards directed again.

Liquid il curve
Hydraulic gradient im, il vs. Line speed vls
0.50 Fixed Bed Cvs=c.
5
Sliding Flow

0.45 Sliding Bed Cvs=c.


Lower Limit
Sliding Bed Cvs=c.
0.40 Mean
8
Hydraulic gradient im, il (m water/m)

Sliding Bed Cvs=c.


Sliding Bed

0.35 Upper Limit


7 Heterogeneous Flow
Cvs=c.
0.30
Equivalent Liquid
Model
0.25
Homogeneous Flow
3 Cvs=Cvt=c.
0.20 Resulting im curve
Cvs=c.
Resulting im curve
0.15 2 Cvt=c.
Limit Deposit Velocity
0.10 Fixed Bed Cvs=c.
Limit Deposit Velocity
0.05 Cvt=c.
1
Limit Deposit Velocity

0.00
0 1 2 3 4 5 6 7 8 9 10
Line speed vls (m/sec)
© S.A.M. Dp=0.1524 m, d=3.000 mm, Rsd=1.585, Cv=0.300, μ=0.416

Figure 5-20: The definition of the pressure losses, scenario’s L3 and R3, im(vls).

Page 148 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Delft Head Loss & Limit Deposit Velocity Framework.

1.1.1.1 Conclusions & Discussion.


From Figure 5-15, Figure 5-16, Figure 5-17, Figure 5-18, Figure 5-19 and Figure 5-20. it is clear that the
characterization of flow regimes of Durand (1952), Abulnaga (2002) or Matousek (2004) is not adequate enough
to identify all possible scenarios. Flow regime graphs like the ones published by Newitt (1955) or King (2002)
(based on Turian & Yuan (1977)) already give a better understanding. These graphs however do not show the
difference between laboratory (Cvs) and real life (Cvt) conditions and do not take the sliding flow effect into
account, probably because the volumetric concentrations were not high enough.

5.2.5. Verification & Validation.


The Relative Excess Hydraulic Gradient Erhg is the contribution of the solids to the relative hydraulic gradient.
The word relative is used here because the hydraulic gradient is divided by the volumetric concentration Cv and
the relative submerged density Rsd in order to determine the Erhg. The Erhg can be applied for all flow regimes.
The relative submerged density Rsd is defined as:

s  l
R sd  (5-16)
l

The Slip Relative Squared Srs is the Slip Velocity of a particle vsl divided by the Terminal Settling Velocity of
a particle vt squared and this Srs value is a good indication of the excess pressure losses due to the kinetic energy
losses of the solids. The Settling Velocity Hindered Relative Shr is the ratio between the hindered settling velocity
vt·(1-Cv/κ)β and the line speed vls, divided by the relative submerged density Rsd and the volumetric concentration
Cv. The Shr value gives a good approximation of the potential energy losses of the solids. The Shr and Srs are
derived and can be applied for the heterogeneous regime.

  
 Cv 
 v  1 
t   2
im  il   C    v sl 
E rh g   R ss  S h r  S rs      (5-17)
R sd  C v  v ls   vt 
 
 

The Stratification Ratio of the Solids R is a measure for the level of stratification of slurry as introduced by
Wilson et al. (1997). A high stratification ratio means that the slurry is (almost) fully stratified; the liquid phase
and the sediment (bed) phase are almost separated.

Under laboratory conditions with constant volumetric spatial concentration Cvs, the Erhg is limited by the value of
the sliding friction coefficient μsf. The lower limit of the Erhg is when the heterogeneous regime transits to the
(pseudo) homogeneous regime. Also here the Srs is derived and can be applied for the heterogeneous regime only.
Resuming, the Erhg is valid for all flow regimes, the Srs is valid for the heterogeneous regime and the friction factor
μsf is valid for the sliding bed regime. In the following examples the Erhg(il) graph will be used. The advantage of
the Erhg(il) graph is that this type of graph is almost independent of the values of concentration Cvs and relative
submerged density Rsd, but also almost independent of the pipe wall roughness and the temperature (kinematic
viscosity). A disadvantage may be that it will take more effort to transform this graph back to real life data, the
hydraulic gradient or pressure versus the line speed.

In these graphs always the lines for fixed/stationary bed (constant Cvs, thick solid red upwards on the left), ELM
(thin dashed dark blue, upwards), homogeneous transport (thick dashed dark brown, upwards), heterogeneous
transport (Cvs thick solid red downwards, Cvt thick dashed green downwards), sliding bed (Cvs thick solid red
horizontal and constant Cvt thick dashed green slightly downwards) are drawn, in order to form a reference system.

The graphs show 3 additional lines. The thin dotted black line shows the ratio of the potential energy to the kinetic
energy for heterogeneous transport. The thin blue dash-dot-dot line shows the heterogeneous curve without
transitions to other flow regimes. The thin brown dash-dot-dot line shows the mobilization of homogeneous
transport. The latter two lines added give the transition from the heterogeneous to the homogeneous flow regime.

All data are compared with the DHLLDV Framework.

Copyright © Dr.ir. S.A. Miedema TOC Page 149 of 304


Introduction Dredging Engineering.

5.2.5.1. L1: Fixed Bed & Heterogeneous, Constant Cvs.

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il


Sliding Bed Cvs=c.

Equivalent Liquid Model


1.000
Relative excess hydraulic gradient Erhg (-)

Homogeneous Flow
Cvs=Cvt=c.
Resulting Erhg curve
Cvs=c.
Resulting Erhg curve
Cvt=c.
Limit Deposit Velocity
0.100

Ratio Potential/Kinetic
Energy
Heterogeneous Flow
with Near Wall Lift
Homogeneous Flow
Mobilized
0.010 Cv=0.170

Cv=0.134

Cv=0.076

Cv=0.036
0.001
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)
© S.A.M. Dp=0.5000 m, d=1.500 mm, Rsd=1.585, Cv=0.036, μsf=0.416
Figure 5-21: Kazanskij (1980), sand, low concentration

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il


Sliding Bed Cvs=c.

Equivalent Liquid Model


1.000
Relative excess hydraulic gradient Erhg (-)

Homogeneous Flow
Cvs=Cvt=c.
Resulting Erhg curve
Cvs=c.
Resulting Erhg curve
Cvt=c.
Limit Deposit Velocity
0.100

Ratio Potential/Kinetic
Energy
Heterogeneous Flow
with Near Wall Lift
Homogeneous Flow
Mobilized
0.010 Cv=0.170

Cv=0.134

Cv=0.076

Cv=0.036
0.001
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)
© S.A.M. Dp=0.5000 m, d=1.500 mm, Rsd=1.585, Cv=0.170, μsf=0.416
Figure 5-22: Kazanskij (1980), sand, high concentration

These experiments clearly show the transition of a fixed bed (flow regimes 1 and 2) to heterogeneous transport
(flow regime 5) at a constant volumetric spatial concentration. The solid lines are drawn for a Cvs=0.036 & 0.17
and may differ slightly for other concentrations. The transition is smoother than the DHLLDV Framework predicts.

Page 150 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Delft Head Loss & Limit Deposit Velocity Framework.

5.2.5.2. R1: Heterogeneous, Constant Cvt.

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il


Sliding Bed Cvs=c.

1.000 Equivalent Liquid Model


Relative excess hydraulic gradient Erhg (-)

Homogeneous Flow
Cvs=Cvt=c.

Resulting Erhg curve


Cvs=c.
0.100 Resulting Erhg curve
Cvt=c.

Limit Deposit Velocity

Ratio Potential/Kinetic
Energy
0.010
Heterogeneous Flow
with Near Wall Lift

Homogeneous Flow
Mobilized

Cv=0.100

0.001
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)
© S.A.M. Dp=0.4400 m, d=0.680 mm, Rsd=1.585, Cv=0.100, μsf=0.416
Figure 5-23: Clift et al. (1982), narrow graded crushed granite.

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il


Sliding Bed Cvs=c.

1.000
Equivalent Liquid
Relative excess hydraulic gradient Erhg (-)

Model

Homogeneous Flow
Cvs=Cvt=c.

Uniform Sand Cvs=c.


0.100

Uniform Sand Cvt=c.

Limit Deposit Velocity

0.010 Graded Sand Cvs=c.

Graded Sand Cvt=c.

Cv=0.110

0.001
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)
© S.A.M. Dp=0.2032 m, d=0.681 mm, Rsd=1.585, Cv=0.110, μsf=0.416
Figure 5-24: Clift et al. (1982), broad graded crushed granite.

These experiments show that at line speeds below the Limit Deposit Velocity and constant Cvt, the heterogeneous
line is still followed. The grading of the sand makes the heterogeneous curve less steep, but this depends on the
grading, the particle size and the pipe diameter.

Copyright © Dr.ir. S.A. Miedema TOC Page 151 of 304


Introduction Dredging Engineering.

5.2.5.3. L2: Fixed & Sliding Bed – Heterogeneous & Sliding Flow, Constant Cvs.

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il


Sliding Bed Cvs=c.

Equivalent Liquid
1.000 Model
Relative excess hydraulic gradient Erhg (-)

Homogeneous Flow
Cvs=Cvt=c.
Limit Deposit Velocity

Graded Sand Cvs=c.

0.100
Graded Sand Cvt=c.

Cv=0.300

Cv=0.250

Cv=0.200
0.010

Cv=0.150

Cv=0.100

Cv=0.050
0.001
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)
© S.A.M. Dp=0.1250 m, d=2.200 mm, Rsd=1.585, Cv=0.150, μsf=0.250
Figure 5-25: Wiedenroth (1967), coarse sand.

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il


Sliding Bed Cvs=c.

Equivalent Liquid
1.000 Model
Relative excess hydraulic gradient Erhg (-)

Homogeneous Flow
Cvs=Cvt=c.
Limit Deposit Velocity

Graded Sand Cvs=c.

0.100
Graded Sand Cvt=c.

Cv=0.300

Cv=0.250

Cv=0.200
0.010

Cv=0.150

Cv=0.100

Cv=0.050
0.001
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)
© S.A.M. Dp=0.1250 m, d=0.900 mm, Rsd=1.585, Cv=0.200, μsf=0.250
Figure 5-26: Wiedenroth (1967), medium sand.

These experiments at constant Cvs, show a fixed bed to sliding bed to heterogeneous behavior. In both figures it is
clear that the curve for graded sands match the data points better. The effect of the grading is different for different
particle sizes.

Page 152 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Delft Head Loss & Limit Deposit Velocity Framework.

1.1.1.2 R2, R3: Sliding Bed & Sliding Flow, Constant Cvt.

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il


Sliding Bed Cvs=c.

Equivalent Liquid Model


1.000
Relative excess hydraulic gradient Erhg (-)

Homogeneous Flow
Cvs=Cvt=c.

Resulting Erhg curve


Cvs=c.

Resulting Erhg curve


0.100 Cvt=c.

Limit Deposit Velocity

Ratio Potential/Kinetic
Energy

Heterogeneous Flow
0.010 with Near Wall Lift

Homogeneous Flow
Mobilized

MnO2, d=1/8 inch


Cv=0.100

MnO2, d=1/16 inch


Cv=0.100
0.001
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)
© S.A.M. Dp=0.0254 m, d=3.000 mm, Rsd=3.000, Cv=0.100, μsf=0.600
Figure 5-27: Newitt et al. (1955), MnO2.

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il


Sliding Bed Cvs=c.

Equivalent Liquid Model


1.000
Relative excess hydraulic gradient Erhg (-)

Homogeneous Flow
Cvs=Cvt=c.
Resulting Erhg curve
Cvs=c.
Resulting Erhg curve
Cvt=c.
0.100
Limit Deposit Velocity

Ratio Potential/Kinetic
Energy
Heterogeneous Flow
with Near Wall Lift
Homogeneous Flow
0.010 Mobilized
Cv=0.210

Cv=0.180

Cv=0.170

0.001
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)
© S.A.M. Dp=0.0508 m, d=3.000 mm, Rsd=0.210, Cv=0.187, μsf=0.416

Figure 5-28: Doron & Barnea (1993), Acetal.

These experiments show the constant Cvt behavior at small line speeds, with a sliding/fixed bed and sliding flow
behavior. The DHLLDV Framework gives a good prediction for both heavy (MnO2) and light (Acetal) solids.

Copyright © Dr.ir. S.A. Miedema TOC Page 153 of 304


Introduction Dredging Engineering.

5.2.5.4. L1, R1, L2, R2:, Homogeneous.

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il


Sliding Bed Cvs=c.

1.000
Relative excess hydraulic gradient Erhg (-)

Equivalent Liquid
Model

Homogeneous Flow
Cvs=Cvt=c.

0.100 Uniform Sand Cvs=c.

Uniform Sand Cvt=c.

Limit Deposit Velocity


0.010

Cv=0.100

Cv=0.100

0.001
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)
© S.A.M. Dp=0.0254 m, d=0.160 mm, Rsd=1.585, Cv=0.100, μsf=0.416

Figure 5-29: Babcock (1970), sand.

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il

Sliding Bed Cvs=c.

1.000
Relative excess hydraulic gradient Erhg (-)

Equivalent Liquid
Model

Homogeneous Flow
Cvs=Cvt=c.
0.100

Uniform Sand Cvs=c.

Uniform Sand Cvt=c.

0.010

Limit Deposit Velocity

Cv=0.240

0.001
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)
© S.A.M. Dp=0.1075 m, d=0.040 mm, Rsd=3.999, Cv=0.240, μsf=0.416

Figure 5-30: Thomas (1976), iron ore.

These experiments show homogeneous behavior, which occurs with small particles at relatively high line speeds.
The Thomas (1976) graph also includes Thomas (1965) viscosity for very small particles.

Page 154 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Delft Head Loss & Limit Deposit Velocity Framework.

5.2.5.5. L3, R3: Sliding Bed & Sliding Flow, Constant Cvs.

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il


Sliding Bed Cvs=c.

Equivalent Liquid Model


1.000
Homogeneous Flow
Relative excess hydraulic gradient Erhg (-)

Cvs=Cvt=c.
Resulting Erhg curve
Cvs=c.
Resulting Erhg curve
Cvt=c.
Limit Deposit Velocity
0.100
Ratio Potential/Kinetic
Energy
Heterogeneous Flow
with Near Wall Lift
Homogeneous Flow
Mobilized
Cv=0.180
0.010
Cv=0.128

Cv=0.107

Cv=0.037

Cv=0.022
0.001
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)
© S.A.M. Dp=0.2000 m, d=4.800 mm, Rsd=1.488, Cv=0.100, μsf=0.470
Figure 5-31: Boothroyde (1979), gravel.

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il


Sliding Bed Cvs=c.

Equivalent Liquid Model


1.000
Relative excess hydraulic gradient Erhg (-)

Homogeneous Flow
Cvs=Cvt=c.
Resulting Erhg curve
Cvs=c.
Resulting Erhg curve
Cvt=c.
Limit Deposit Velocity
0.100

Ratio Potential/Kinetic
Energy
Heterogeneous Flow
with Near Wall Lift
Homogeneous Flow
Mobilized
0.010 Cv=0.200

Cv=0.150

Cv=0.100

Cv=0.050
0.001
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)
© S.A.M. Dp=0.1250 m, d=5.950 mm, Rsd=1.585, Cv=0.150, μsf=0.416
Figure 5-32: Wiedenroth (1967), gravel.

These experiments show the sliding bed and sliding flow regimes. Beyond the intersection point between a sliding
bed and heterogeneous flow, the sliding flow curve is followed.

Copyright © Dr.ir. S.A. Miedema TOC Page 155 of 304


Introduction Dredging Engineering.

5.2.6. Discussion & Conclusions.


The experimental graphs are given without a lot of explanation, because they should speak for themselves. These
graphs show the different flow regimes and sometimes more than one flow regime. In general there is a lot of
scatter. This is caused by the way experiments were carried out and specifically the accuracy of the concentration
measurements. Sometimes concentrations within a certain bandwidth (for example 10-15%) are given with an
average mentioned on the graph (for example 12.5%). But in spite of the scatter, the graphs clearly show the
different regimes.

From these graphs and the regime and scenario definitions, it should be clear that experiments carried out in very
small pipelines, like 1 inch diameter pipelines, can never be compared with experiments in very large pipelines,
like 1 m diameter pipelines. In a 1 m diameter pipeline it is difficult to get a sliding bed regime, while in a 1 inch
diameter pipeline it is very difficult not to get a sliding bed regime, due to the high hydraulic gradients. It is like
comparing laminar and turbulent flow.

Each regime has its own physical and mathematical model. The fixed bed regime can be modeled with flow
through a restricted cross section using the Televantos (1979) method for determining the friction factor. The
sliding bed regime and partly the sliding flow regime can be modeled using the Newitt et al. (1955) method, with
the appropriate friction factor (0.35-0.7). The heterogeneous regime can be modeled with one of the existing
equations or with the Miedema et al. (2013) model. The homogeneous regime can be modeled using the equivalent
liquid model, using 100% of the solids or for example using 60% of the solids, like some authors do. For the fixed
bed/sliding bed regimes below the Limit Deposit Velocity, a 2 layer or 3 layer model can be used, but the Durand
& Condolios (1952) approach, considering a flow Froude number which is equal to the flow Froude number at the
Limit Deposit Velocity, also gives good results.

Just carrying out some curve fits and drawing conclusions is very dangerous, because the experiments may cover
2 or more regimes. For example, if 50% of the experiments are in the heterogeneous regime and 50% of the
experiments are in the homogeneous regime, a curve fit would give a horizontal line in the Erhg(il) graph. If we
look at experiments where 50% is in the fixed bed regime (constant Cvs) and 50% is in the heterogeneous regime
(for example Figure 5-21 & Figure 5-25), the result of a curve fit is also a horizontal line. The cases however are
completely different.

Recognizing the different regimes and especially the transitions between the different regimes is crucial in
understanding what is physically happening.

The DHLLDV Framework gives a good match with most of the experiments shown here. In this book many more
experiments are shown, usually compared to the DHLLDV Framework. The DHLLDV Framework gives good
results for fine sands up to fine gravel in small to large pipes, compared to the experiments. For very coarse
particles however still more research is required.

Page 156 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Delft Head Loss & Limit Deposit Velocity Framework.

5.2.7. Nomenclature Flow Regimes & Scenario’s.

A2 Cross section of the bed in the pipe m2


A1 Cross section of the mixture in suspension above the bed m2
A, Ap Cross section of the pipe m2
A2 Cross section of the solids in the pipe m2
Cv Volumetric concentration -
Cvs Volumetric spatial concentration -
Cvs,s Volumetric spatial concentration of the mixture in suspension above a fixed or
sliding bed -
Cvt Volumetric transport (delivered) concentration -
d Particle diameter mm
d50 Median particle diameter - 50% by weight is smaller mm
Dp Pipe diameter m
Erhg Relative Excess Hydraulic Gradient -
im Mixture head loss m/m
il Liquid hydraulic gradient m/m
iw Water hydraulic gradient m/m
n Porosity -
O1 Circumference liquid-pipe wall m
O2 Circumference bed-pipe wall m
O12 Width liquid-bed interface m
Qm or Vm Flow rate of mixture m3/s
Qs or Vs Flow rate of solids m3/s
Rsd Relative submerged density -
R Stratification ratio Wilson -
Srs Slip Relative Squared or Stratification Ratio Solids -
Shr Settling velocity Hindered Relative -
vls Velocity of the slurry, line speed m/s
vls,e Effective line speed. Line speed above the fixed or moving bed. m/s
vm Velocity of the slurry, line speed (same as vls) m/s
vs Average velocity of the solids m/s
vsl Slip velocity of the solids relative to the mixture m/s
Vb Volume of the bed m3
Vcl Volume of the closed loop m3
Vm Volume of the mixture in a pipe m3
Vm,s Volume of the mixture in suspension above the bed m3
Vs Volume of the solids in a pipe m3
Vs,s Volume of the solids in suspension above the bed m3
vt Terminal settling velocity of the particles m/s
β Richardson & Zaki hindered settling power -
λ1 Darcy-Weisbach friction factor between liquid and pipe wall -
λ2 Darcy-Weisbach friction factor between liquid and pipe wall in bed -
λ22 Darcy-Weisbach friction factor between liquid and bed -
κC Concentration eccentricity factor. -
μsf Friction coefficient for a sliding bed -
ρl, ρw Density of the liquid ton/ m3
ρm Density of the mixture ton/ m3
ρs Density of the solids ton/ m3

Copyright © Dr.ir. S.A. Miedema TOC Page 157 of 304


Introduction Dredging Engineering.

Page 158 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Delft Head Loss & Limit Deposit Velocity Framework.

5.3. Notes.

Copyright © Dr.ir. S.A. Miedema TOC Page 159 of 304


Introduction Dredging Engineering.

Page 160 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Delft Head Loss & Limit Deposit Velocity Framework.

Copyright © Dr.ir. S.A. Miedema TOC Page 161 of 304


Introduction Dredging Engineering.

Page 162 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Delft Head Loss & Limit Deposit Velocity Framework.

Copyright © Dr.ir. S.A. Miedema TOC Page 163 of 304


Introduction Dredging Engineering.

Page 164 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Slurry Transport Models.

Chapter 6: Slurry Transport Models.

6.1. The DHLLDV Framework.


6.1.1. The Sliding Bed Regime.
For the sliding bed the same set of equations is used as for the stationary bed. The only difference is that for a
stationary bed the bed velocity v2 equals 0, while of course for a sliding bed the bed has a positive velocity v2,
smaller than v1. A convenient parameter to show the results of the calculations is the Relative Excess Hydraulic
Gradient. This parameter gives an almost dimensionless graph of the head losses.

im  il
E rh g  (6-1)
R sd C vs

Figure 6-1 shows the Erhg parameter as a function of the relative volumetric concentration (Cvr=Cvs/Cb) and the
relative line speed (vls/vls,ldv,max) for the weight approach sliding bed friction (Miedema & Ramsdell (2014)) and a
sliding bed friction factor μsf=0.4. The Erhg parameter is very close to the sliding friction coefficient μsf, especially
for relative line speeds up to 1.5, the region where most probably the sliding bed will occur.

So for the sliding bed regime the Erhg parameter is defined to be equal to the sliding friction coefficient μsf.

im  il
E rh g    sf (6-2)
R sd  C vs

2LM-M-W
Erhg value vs. relative line speed vr, Cvr=Cvs/Cvb
1.00
LSDV

0.90
Cvr=0.05

0.80 Cvr=0.10

Cvr=0.20
Erhg=(im-il)/(Rsd·Cvs) (-)

0.70

Cvr=0.30
0.60

Cvr=0.40
0.50
Cvr=0.50

0.40
Cvr=0.60

0.30 Cvr=0.70

0.20 Stationary Cvr=0.80

Deposit
Cvr=0.90
0.10

Cvr=0.95
0.00
0.0 0.5 1.0 1.5 2.0 2.5
Relative line speed vr=vls/vsm (-)
© S.A.M. Dp=0.7620 m, d=1.00 mm, Rsd=1.59, μsf=0.415, Cvb=0.55

Figure 6-1: The Erhg parameter versus the relative line speed.

The sliding friction coefficient μsf is the tangent of the external friction angle δ between the sand or gravel and the
steel pipe wall. From soil mechanics it is known that the external friction angle δ is about 2/3 of the internal friction
angle φ. This internal friction angle has a minimum of about 30° for loose packed sand, giving 20° for the external
friction angle. The tangent of 20° is 0.364. Miedema & Ramsdell (2014) also analyzed the hydrostatic approach
of Wilson et al. (1992) and the normal stress carrying the weight approach. These two approaches are similar up
to a relative concentration of 0.5, giving an increase of the Erhg parameter with a factor 1.3 compared to the weight
approach as used here. In practice the relative concentration will be between 0 and 0.5 giving a multiplication

Copyright © Dr.ir. S.A. Miedema TOC Page 165 of 304


Introduction Dredging Engineering.

factor between 1 and 1.3 depending on the relative concentration. Taking an average gives a sliding friction factor
of about 0.415. Resuming it can be stated that the Erhg parameter should have a value of about 0.364 if the weight
approach is applied, or a value of about 0.415 if the hydrostatic or normal stress carrying the weight approach are
applied. In the current model a constant value of 0.415 is used to be on the safe side, resulting in hydraulic gradient
curves parallel to the liquid curve as already observed by Newitt et al. (1955) and others.

6.1.2. The Heterogeneous Transport Regime.


Miedema & Ramsdell (2013) derived an equation for the Relative Excess Hydraulic Gradient for heterogeneous
transport based on energy considerations. This equation consists of two parts. A first part for the contribution due
to potential energy losses and a second part for the kinetic energy losses. The equation is based on uniform sands
or gravels, but Miedema (2014) also derived a modified equation for graded sands and gravels. In its basic form
the equation looks like:


 C vs 
vt 1   2
im  il  C   v sl  (6-3)
E rh g       S h r  S rs
R sd
C vs
v ls  vt 

The Settling Velocity Hindered Relative, Shr, is the Hindered Settling Velocity of a particle vt·(1-Cvs/κ)β divided
by the line speed vls. The Shr value gives the contribution of the potential energy losses to the Relative Excess
Hydraulic Gradient. The Shr is derived for and can be applied to the heterogeneous regime.
The Slip Relative Squared Srs is the Slip Velocity of a particle vsl divided by the Terminal Settling Velocity of
a particle vt squared and this Srs value is a good indication of the Relative Excess Hydraulic Gradient due to the
solids, since its contribution to the total is 90%-100%. The Srs value gives the contribution of the kinetic energy
losses to the Relative Excess Hydraulic Gradient. The Srs is derived for and can be applied to the heterogeneous
regime.

The potential energy term is explicit and all the variables involved are known, so this term can be solved. The
kinetic energy term however contains the slip velocity, which is not known. The kinetic energy term has been
derived by Miedema & Ramsdell (2013) based on kinetic energy losses due to collisions or interactions with the
pipe wall or the viscous sub layer. This means that the slip velocity used in the above equation is not necessarily
the average slip velocity, but it is the slip velocity necessary to explain the kinetic energy losses. The average slip
velocity of the particles will probably be larger, but of the same magnitude. The derivation of the slip velocity
equation for uniform sands or gravels will be subject of another paper, but the resulting equation for the Erhg
parameter is given here. Giving for the relative excess hydraulic gradient, the Erhg parameter:


 C vs 
vt 1   10 / 3 2
   g 1/3 
im  il  C   1   vt   l 
 
2
E rh g   S h r  S rs   8 .5    
R sd  C vs  g  d   
v ls  l   
v ls

(6-4)

 C  0 .1 7 5   1   

The equation has been modified slightly since the original article of Miedema & Ramsdell (2013). The derivation
is published in Miedema (2015).

6.1.3. The Homogeneous Transport Regime.


The basis of the homogeneous transport regime model is the equivalent liquid model (ELM). In terms of the
relative excess hydraulic gradient, Erhg, this can be written as:

2
im  il  l  v ls
E rh g    il (6-5)
R sd  C vs 2 g Dp

Talmon (2013) derived an equation to correct the homogeneous equation (the ELM model) for the slurry density,
based on the hypothesis that the viscous sub-layer hardly contains solids at very high line speeds in the

Page 166 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Slurry Transport Models.

homogeneous regime. This theory results in a reduction of the resistance compared with the ELM, but the
resistance is still higher than the resistance of clear water. Talmon (2013) used the Prandl approach for the mixing
length, which is a 2D approach for open channel flow with a free surface. The Prandl approach was extended with
damping near the wall to take into account the viscous effects near the wall, according to von Driest (Schlichting,
1968). Miedema (2015) extended the model with pipe flow and a concentration distribution, resulting in the
following equations.

The value of the Darcy Weisbach wall friction factor l depends on the Reynolds number:

v ls  D p
Re  (6-6)
l

Over the whole range of Reynolds numbers above 2320 the Swamee Jain (1976) equation gives a good
approximation:

1 .3 2 5 0 .2 5
l  
2 2
   5 .7 5     5 .7 5  (6-7)
 ln     lo g 1 0   
  3 .7  D Re
0 .9    3 .7  D Re
0 .9 
  p    p 

For the resulting Darcy-Weisbach friction factor ratio this can be approximated by:

m 1

l  AC 
2
 m  l (6-8)
 ln    1
v

  
  l  8

The hydraulic gradient il (for a liquid including the fine solids effect in general) is:

2
pl  l  v ls
il   (6-9)
l  g  L 2 g Dp

The relative excess hydraulic gradient Erhg is now:

2
 AC  m  l 
1 R Cv    ln    1
v
sd   
im  il   l  8

E rh g   il    E  il
R  C vs  AC 
2 (6-10)
sd  m  l
Cv   ln    1
v
R sd   
  l  8

Table 6-1: Some A and ACv values.

A ACv
Law of the Wall 4.13 1.00
Nikuradse (no damping) 3.02 1.25
Prandtl (damping) 0.01 3.40
Average 1.05 2.20
Lower limit of data 5.43 0.80
Upper limit of data 1.67 1.80

A value of ACv=3 is adviced.

Copyright © Dr.ir. S.A. Miedema TOC Page 167 of 304


Introduction Dredging Engineering.

6.1.4. The Resulting Erhg Constant Spatial Volumetric Concentration Curve.


The hydraulic gradient for the mixture can be determined with:

i m  i l  E rh g  R sd C vs (6-11)

and

p m  i l   l  g  L  E rh g  R sd  C vs   l  g  L  p l  p s (6-12)

6.1.5. Determining the Limit Deposit Velocity.

6.1.5.1. Introduction.

Figure 6-2: The algorithm to determine the Limit Deposit Velocity.

Figure 6-2 shows the algorithm to determine the Limit Deposit Velocity. The different steps are discussed in the
next chapters.

Page 168 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Slurry Transport Models.

6.1.5.2. Very Small & Small Particles.


For very small particles, smaller than about 50% of the thickness of the viscous sub layer, the LDV and the Froude
number FL are:

1/3 8
v ls , ld v  1 .4    l  R sd g 
l

1/3 8 (6-13)
1 .4    l  R sd g 
v ls , ld v l
FL , v s  
1/2 1/2
 2  g  R sd  D p   2  g  R sd  D p 

For small particles and a smooth bed, in the case of sand particles smaller than about d=0.15 mm, this gives for
the Limit Deposit Velocity:


 C vs 
vt 1 
C 
 
 C vs  2  g  R sd  D p  (6-14)
3 3 
v ls , ld v  p 
l

In terms of the Durand & Condolios LDV Froude number FL factor this can be written as:

1/3
  
 C vs 
 v 1  C 
t  vs
v ls , ld v   C  
F L ,s s   p  
1/2 1/2
 2  g  R sd  D p    

l  2  g  R sd  D p  
 (6-15)
 
2/9
 1 .6 5 
w it h :  p  3 .4    ,  C = 0 .1 7 5   1   
 R sd 

The coefficient αp=3.5 is an upper limit. The minimum found is about 3.0, the average 3.25. To be on the safe
side, the value of 3.5 should be used. To find the highest correlation with experimental data, the value of 3.25
should be used. With the following conditions the Froude number FL for very small and small particles can be
determined:

If F L , v s  F L ,s s  F L ,s  F L , v s
(6-16)
If F L , v s  F L ,s s  F L ,s  F L ,s s

6.1.5.3. Large & Very Large Particles.


The Limit Deposit Velocity LDV is for medium and large particles and a rough bed:

 1/2
 C vs     1/2
1   C vs    sf  C vb    C v r , ld v
 C   8  3/2 (6-17)
3
v ls , ld v   p 
3

l

 2  g  R sd  D p 

And the Durand & Condolios LDV Froude number:

Copyright © Dr.ir. S.A. Miedema TOC Page 169 of 304


Introduction Dredging Engineering.

1/3
  1/2 
 C vs    
 1 1/2 
   C vs    sf  C vb    C v r , ld v
v ls , ld v   C   8  
F L ,r   p   (6-18)
1/2 l
 2gR sd Dp  



 

The bed fraction at the Limit Deposit Velocity is, depending on the particle diameter to pipe diameter ratio:

1
1  R sd  0 .0 0 5 5
d  0 .0 1 5  D p C v r , ld v  0 .0 0 0 1 7  D p   
 1 .6 5  2g R sd Dp

(6-19)
1 1/2 1/2
 R sd   d  0 .0 4 5  d 
1
d  0 .0 1 5  D p C  0 .0 0 0 1 7  D p       
v r , ld v
 1 .6 5   0 .0 1 5  D  2g R Dp  D 
 p  sd  p 

6.1.5.4. The Resulting Upper Limit Froude Number.


The resulting upper limit of the Froude number FL,ul value can now be determined according to (for sand):

F L ,s  F L ,r  F L ,u l  F L ,s

F L ,s  F L ,r  F L ,u l  F L ,s  e
 d /d0
 F L ,r  1  e   d /d 0

(6-20)
d  d rou gh  F L ,u l  F L ,r

2
 1 .6 5 
w it h : d 0  0 .0 0 0 5   
 R sd 

The value of 2 mm is valid for sand with Rsd≈1.65, other materials will have a different value. This value will
increase with decreasing solids density.

6.1.5.5. The Lower Limit.


The lower limit of the LDV is the transition velocity between the sliding bed regime and the heterogeneous regime,
resulting in the transition velocity at:

 10 / 3
 C vs  8 .5
2  1  2/3
vt 1    v ls , ld v    l g
C  l   (6-21)
2   Cx 
v ls , ld v 
 sf

This equation shows that the transition between the sliding bed regime and the heterogeneous regime depends on
the sliding friction coefficient. The equation derived is a second degree function and can be written as:

10 / 3
 2  
 C vs  8 .5 1 2/3
vt 1     l g
C  l  Cx  (6-22)
2   
 v ls , ld v   v ls , ld v   0
 sf  sf

With:

Page 170 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Slurry Transport Models.

A  1


 C vs 
vt 1  
 C 
B 
 sf

10 / 3 (6-23)
8 .5
2  1  2/3
  l g
l  Cx 
 
C 
 sf

2
B  B  4A C
v ls , ld v 
2A

In terms of the Durand & Condolios LDV Froude number FL factor this can be written as:

v ls , ld v
F L , ll 
1/2 (6-24)
 2  g  R sd  D p 

Durand Froude number FL (-) vs. Particle diameter d (m)


Particle diameter d (mm)
0.0 0.5 1.0 1.5 2.0 2.5 3.0
2.0
FL Min
1.8 FL Max
FL,ul=FL,s·e-d/d0+FL,r·(1-e-d/d0) FL Cvs=0.01
1.6 FL Cvs=0.02
Durand Froude number FL (-)

FL,ul=FL,r FL Cvs=0.03
1.4
FL Cvs=0.04

1.2
FL Cvs=0.05
FL,ll FL Cvs=0.07
1.0 FL Cvs=0.10
FL Cvs=0.15
0.8 FL,ul=FL,ss FL Cvs=0.20
FL Cvs=0.25
0.6
Cvs=2%
0.4 Cvs=5%
Cvs=10%
0.2 FL,ul=FL,vs
Cvs=15%
Cvs=20%
0.0
0.0000 0.0005 0.0010 0.0015 0.0020 0.0025 0.0030
Particle diameter d (m)
© S.A.M. Dp=0.1524 m, Rsd=1.59, Cvs=0.175, μsf=0.550

Figure 6-3: The resulting FLcurves.

Copyright © Dr.ir. S.A. Miedema TOC Page 171 of 304


Introduction Dredging Engineering.

6.1.5.6. The Resulting Froude Number.


The resulting Froude FL value can now be determined according to:

F L ,u l  F L , ll  F L  F L ,u l

(6-25)
F L , ll  F L ,u l  F L  F L , ll

For small particles and/or concentrations near 20% and/or large pipe diameters, usually the upper limit Froude
number will be valid. For large particles and/or low concentrations and/or very small to small pipe diameters,
usually the lower limit Froude number will be valid.

Figure 6-3 shows the resulting LDV curves for a number of volumetric concentrations, including the Durand &
Condolios (1952) data. The graph matches the graph as published by Durand (1953) very well. The use of the
lower limit based on the transition sliding bed regime to heterogeneous regime is not exact, since this transition
velocity will not be exact. It is possible that this lower limit should be set to 90% or 95% of this transition velocity.
Here the theoretical transition velocity is used.

Page 172 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Slurry Transport Models.

6.1.6. Nomenclature DHLLDV Framework.

ACv Coefficient homogeneous regime (1.3 by default) -


Ap Cross section of the pipe m2
Cvb Bed volumetric concentration -
Cvb,max Maximum bed volumetric concentration -
CvB Concentration at the bottom of the pipe -
CD Particle drag coefficient -
Cvs Spatial volumetric concentration -
Cvr Relative concentration Cvs/Cvb -
Cvr,ldv Relative concentration in bed at LDV -
Cvt Transport or delivered volumetric concentration -
Cx Durand & Condolios coefficient -
d Particle diameter m
d0 Particle diameter LDV transition region m
DH Hydraulic diameter m
Dp Pipe diameter m
Erhg Relative excess hydraulic gradient -
Erhg,SF Relative excess hydraulic gradient in the sliding flow regime -
Erhg,HeHo Relative excess hydraulic gradient in the heterogeneous/ homogeneous flow regimes -
f Fraction of fines -
f Factor determining sliding flow -
FL Durand limit deposit velocity Froude number -
FL,s Durand limit deposit velocity Froude number, smooth bed -
FL,ss Durand limit deposit velocity Froude number, small particles smooth bed -
FL,vs Durand limit deposit velocity Froude number, smooth bed, very small particles -
FL,r Durand limit deposit velocity Froude number, rough bed, large particles -
FL,ul Durand limit deposit velocity Froude number, upper limit -
FL,ll Durand limit deposit velocity Froude number, lower limit -
FrDC Durand & Condolios Froude number -
g Gravitational constant (9.81) m/s2
h Thickness of bed at LDV m
il Hydraulic gradient of liquid -
im Hydraulic gradient of mixture -
im,i Hydraulic gradient of ith fraction of PSD -
im,ldv Hydraulic gradient mixture at LDV -
K Durand & Condolios constant (85) -
mp Mass particle kg
N Zandi & Govatos deposit criterion -
r Position in pipe starting at the bottom -
Re Reynolds number based on velocity difference liquid flow - bed -
Rsd Relative submerged density -
Shr Settling Velocity Hindered Relative -
Srs Slip Velocity Relative Squared -
u* Friction velocity m/s
u*,ldv Friction velocity at the LDV m/s
v1 Average velocity above the bed m/s
v2 Velocity of the bed m/s
v12 Velocity difference bed interface (v1-v2) m/s
vls Cross-section averaged line speed m/s
vls,ldv Limit Deposit Velocity (LDV) m/s
vr Relative line speed vls/vls,ldv,max or vls/vsm m/s
vsl Slip velocity (velocity difference between particle and liquid) m/s
vsm Maximum LSDV according to Wilson m/s
vt Particle terminal settling velocity m/s
vth Hindered settling velocity m/s
vtv (Hindered) settling velocity in the vehicle (Wasp model) m/s
vtv,ldv (Hindered) settling velocity in the vehicle (Wasp model) at LDV m/s

Copyright © Dr.ir. S.A. Miedema TOC Page 173 of 304


Introduction Dredging Engineering.

αh Coefficient homogeneous equation -


αp LDV factor -
αsm Factor concentration distribution -
β Angle of bed with vertical rad
β Power of Richardson & Zaki hindered settling factor -
βsm Relation sediment diffusivity eddy momentum diffusivity -
φ Internal friction angle rad
δ External friction angle rad
λ 1, λ l Darcy Weisbach friction factor liquid to pipe wall -
λ12 Darcy Weisbach friction factor bed interface -
κ Von Karman constant (about 0.4) -
κC Concentration distribution constant -
ρl Density of liquid ton/m3
ρl,m Density of liquid including fines ton/m3
ρm Mixture density ton/m3
ρs Density of solids ton/m3
νl Kinematic viscosity liquid m2/s
μl Dynamic viscosity liquid Pa·s
μl,m Dynamic viscously liquid including fines Pa·s
μsf Sliding friction coefficient -
τ2,sf Shear stress bed – pipe wall due to sliding friction kPa
τ12 Bed shear stress kPa
ξ Slip ratio -
ξ0 Slip ratio asymptotically for line speed zero -
ζ Bed fraction -

Page 174 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Slurry Transport Models.

6.2. The Jufin & Lopatin (1966) Model.


6.2.1. Introduction.
The Jufin & Lopatin (1966) model was constructed as a proposal for the Soviet technical norm in 1966. The authors
did not submit a new model but selected the best combination of correlations for the frictional head loss and the
critical velocity from four models submitted by different Soviet research institutes. The four models submitted
were tested by a large experimental database collected by a number of researchers. The database contained data
from both laboratory and field measurements (including data from dredging installations). The data covered a wide
range of pipeline sizes (24 – 900 mm) and particle sizes (sand and gravel, 0.25 - 11 mm). Some of the data on
which the model is based can be found in the chapters about Silin, Kobernik & Asaulenko (1958), Kazanskij (1978)
and on the website www.dhlldv.com.
Kazanskij (1972) gave a summary and sort of manual for the use of the Jufin-Lopatin model. First of all sands and
gravels are divided into 4 groups, according to Table 6-2. The ψ* parameter characterizes the particles and is
comparable with the Durand & Condolios (1952) √Cx parameter.

Table 6-2: Group classification of Jufin-Lopatin (1966), source Kazanskij (1972).


Group Range ψ*
A d<0.06 mm -
B d60<10 mm All
d10<10 mm<d60 ψ*<=1.5, d0<2.5 mm
C d10<10 mm<d60 ψ*>1.5, d0>2.5 mm
D d10>10 mm -

The particle diameter d0 is the average particle diameter, not a weighted particle diameter, and can be determined
by:

100 90

 di  di
(6-26)
i1 i10
d0  or d0 
100 9

So each fraction has the same weight in the determination of the d0 value. For uniform sands and gravels, the d0 is
equal to the particle diameter.

6.2.2. Group A: Fines.


Group A covers the fines, silt. For silt Jufin & Lopatin (1966) use the ELM without the Thomas (1965) viscosity,
so:

L 1 2
pm   l     m  v ls (6-27)
Dp 2

The hydraulic gradient im (for mixture) is now:

2
pm m  l  v ls
im    (6-28)
l  g  L l 2 g  Dp

6.2.3. Group B: Sand.


Group B covers fine and medium sands, with possibly some fine gravel. The equation found by Jufin & Lopatin
(1966) was based on the empirical experience, suggesting that the minimum hydraulic gradient at the velocity vmin
was independent of the mixture flow properties and it was 3 times higher than the hydraulic gradient of water flow
at the same velocity in a pipeline. This was also experienced in the American dredging industry (see Turner
(1996)). Now most frictional head models follow the equation:

Copyright © Dr.ir. S.A. Miedema TOC Page 175 of 304


Introduction Dredging Engineering.

  1  
3

pm  pl 1  2    


1/3
 w ith : v m in   (6-29)
  v ls  

The minimum is found at the cube root of Ω, as is the case with the Durand/Condolios/Gibert, Newitt et al. and
Fuhrboter models. The frictional-head-loss correlation by Jufin & Lopatin is:

 3 
 v m in 
pm  pl 1  2  
 (6-30)
  v ls  
 

With, for quarts particles (sometimes a factor 5.3 is used instead of 5.5) :

     
1/6 1/6 1/6
* * *
v m i n  5 .5  C vt   Dp  3 .7 6  C vt  g Dp  3 .4 6  C vt  g Dp R sd (6-31)

And:

3/2
 vt 
* 3/2
  F rp    (6-32)
 g  d 

This can be written in a more general form for the hydraulic gradient according to:

 3 
 1 
 
1/2
i m  i l   1  2  4 1 .3 5  C v t    g  D p  R s d 
*
  (6-33)
  v ls  
 

Jufin-Lopatin ψ* compared with Gibert & DHLLDV


2.50
Jufin-Lopatin

2.25 Jufin-Lopatin Data

Durand, Condolios & Gibert


2.00
H2 Sand, d50=0.20 mm,
√(Cx)=3.42
1.75 L3 Sand, d50=0.27 mm,
√(Cx)=1.96
L4 Sand, d50=0.37 mm,
1.50 √(Cx)=1.34
ψ* (-)

L5 Sand, d50=0.58 mm,


√(Cx)=1.06
1.25
L6 Sand, d50=0.89 mm,
√(Cx)=0.88
1.00
L7 Sand, d50=1.33 mm,
√(Cx)=0.80
L8 Sand, d50=2.05 mm,
0.75 √(Cx)=0.72
A9 Gravel, d50=2.80 mm,
√(Cx)=0.67
0.50 A10 Gravel, d50=4.20 mm,
√(Cx)=0.62
Theory ψ*=Frp^1.5
0.25
DHLLDV & Gibert
ψ*=Frp^10/3
0.00 Wilson NWL
0.1 1 10
d (mm) (Gibert+Wilson)/2
© S.A.M.

Figure 6-4: The Jufin-Lopatin ψ* compared to Gibert and DHLLDV.

Figure 6-4 shows the ψ* parameter of Jufin & Lopatin (1966) according to Kazanskij (1972). This parameter is
compared with the equivalent parameters of Gibert (1960), Fuhrboter (1961) and the DHLLDV framework
(Miedema S. A., 2014). The trends are similar, but especially for medium sands, the values differ.

Page 176 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Slurry Transport Models.

Figure 6-5 shows a comparison in terms of the Gibert (1960) √Cx value. Of course the trends are similar compared
to Figure 6-4.

Reciprocal Particle Froude Number √Cx Gibert & DHLLDV


5
Jufin-Lopatin

Jufin-Lopatin data

Durand, Condolios & Gibert

H2 Sand, d50=0.20 mm,


√(Cx)=3.42
L3 Sand, d50=0.27 mm,
√(Cx,Gibert) or ψ*-2/3 (-)

√(Cx)=1.96
L4 Sand, d50=0.37 mm,
√(Cx)=1.34
L5 Sand, d50=0.58 mm,
√(Cx)=1.06
L6 Sand, d50=0.89 mm,
√(Cx)=0.88
L7 Sand, d50=1.33 mm,
√(Cx)=0.80
L8 Sand, d50=2.05 mm,
√(Cx)=0.72
A9 Gravel, d50=2.80 mm,
√(Cx)=0.67
A10 Gravel, d50=4.20 mm,
√(Cx)=0.62
Theory √(Cx)=1/Frp

DHLLDV & Gibert


√(Cx)=1/Frp^20/9
0.5 Wilson NWL
0.1 1 10
d (mm) (Gibert+Wilson)/2
© S.A.M.

Figure 6-5: The reciprocal particle Froude number of Jufin-Lopatin, Gibert and DHLLDV.

Assuming the experiments are carried out with quarts this can be written as:

 3/4 3 
 vt  1/2  1 
i m  i l   1  2  4 1 .3 5     
1/2
  g  D p  R sd   C vt    (6-34)
  g  d  
  v ls  
 

The term vmin should have the dimension of velocity, but in equation (6-31) it has the dimension of the cube root
of velocity. This has to be compensated without violating the model of Jufin Lopatin. Now the product of kinematic
viscosity ν and the gravitational constant g has the dimension of velocity to the 3rd power. It is not clear whether
Jufin & Lopatin carried out experiments in liquids with different viscosities, but for dredging purposes it is neutral
using this. This gives for vmin, using a kinematic viscosity of 10-6 m2/sec and a gravitational constant of 9.81 m/sec2:

 
1/6 2/9
l g
*
v m in  4 4 .8 8  C v t    g  D p  R s d (6-35)

Substituting this in equation (6-30) gives the equation for the hydraulic gradient.

 3/4 3 
 vt  1/2  1 
 il   1  2  90389   
2/3 1/2
im    g  D p  R sd l g   C vt    (6-36)
  g  d  
  v ls 
 

Copyright © Dr.ir. S.A. Miedema TOC Page 177 of 304


Introduction Dredging Engineering.

6.2.4. The Limit Deposit Velocity.


Jufin & Lopatin (1966) defined the Limit Deposit Velocity as (sometimes a value of 8 is used instead of 8.3):

 
1/6
* 1/3
v ls , ld v  8 .3  C v t    Dp (6-37)

It is clear that this Limit Deposit Velocity also does not have the dimension of velocity, but the cube root of length.
To give this Limit Deposit Velocity the dimension of velocity, the equation is modified to (for quarts and water):

 
1/6 1/3
 
1/9
l g
*
v ls , ld v  9 .2 3  C v t    2  g  D p  R sd (6-38)

Which can be written as:

1/4
 vt  1/3
 
1/6 1/9
v ls ,  9 .2 3   C     2 g Dp R l g  (6-39)
ld v vt  g  d 
sd

Giving:

1/4
1/6  vt  1/9
 C vt  

 l g 
v ls , ld v  g  d  (6-40)
FL   9 .2 3 
1/2 1/6
 2  g  D p  R sd   2  g  D p  R sd 

6.2.5. Broad Graded Sands or Gravels.


The effect of a broad particle size distribution is taken into account by determining an average value of the modified
particle Froude number from values of the modified Froude number for soil fraction pi of different size di. The
values for ψ* can also be taken from Table 6-2 or Figure 6-4.

n n

 
* 1 .5 1 .5 *
  F rv t  F r v t ,i  p i   (d i )  p i (6-41)
i1 i1

Table 6.1: Particle settling parameter for the Jufin-Lopatin (1966) model.
particle settling
particle settling
size fraction of solids, parameter, ψ*
parameter, ψ*
d [mm] Jufin & Lopatin
Jufin (1971)
(1966)

0.05 - 0.10 0.0204 0.02


0.10 - 0.25 0.0980 0.2
0.25 - 0.50 0.4040 0.4
0.50 - 1.00 0.7550 0.8
1.0 - 2.0 1.1550 1.2
2.0 - 3.0 1.5000 1.5
3.0 - 5.0 1.7700 1.8
5 - 10 1.9400 1.9
10 - 20 1.9700 2.0
>20 2.0000 2.0

Page 178 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Slurry Transport Models.

6.2.6. Group C: Fine Gravel.


Group C is a transition between medium sized sand and coarse gravel. The equation for vmin has to be corrected
according to:

   
1/6 1/6
* *
v m in  5 .5  b  C vt   Dp  3 .7 6  b  C vt  g  Dp

(6-42)

   
1/6 1/6 2/9
l g
* *
 3 .4 6  b  C vt  g Dp R sd  4 4 .8 8  b  C vt  g  Dp  R sd

Table 6-3: Correction factor a, source Kazanskij (1972) .


d0 10 mm<d0<20 mm d0>20 mm
ρm (ton/m3) 1.02 1.05 1.10 1.20 1.02 1.05 1.10 1.20
Dp<400 mm 1.01 1.18 1.34 1.48 1.11 1.30 1.48 1.68
400 mm<Dp<600 mm 1.14 1.31 1.47 1.64 1.27 1.46 1.62 1.81
600 mm<Dp 1.23 1.41 1.54 1.73 1.38 1.50 1.67 1.86

The correction factor b can be determined with:

  1 .5 
b  1  a  1 (6-43)
 2 .0  1 .5 

Where the factor a can be found in Table 6-3.

6.2.7. Group D: Coarse Gravel.


For Group D the correction factor is just a, according to Table 6-3.

1/6 1/6

v m in  5 .5  a  C v t  2  D p  
 3 .7 6  a  C v t  2  g  D p 
(6-44)
1/6 1/6
   
2/9
 3 .4 6  a  C v t  2  g  D p  R s d  4 4 .8 8  a  C v t  2  g  D p  R sd l g

Copyright © Dr.ir. S.A. Miedema TOC Page 179 of 304


Introduction Dredging Engineering.

6.2.8. Conclusions & Discussion.


The model of Jufin & Lopatin (1966) for Group A is the ELM model without a viscosity correction.
The models for Groups B, C and D are similar, but Groups C and D have a correction factor. In order to make the
Jufin & Lopatin (1966) model comparable with other models, the basic equation is written in terms of the liquid
hydraulic gradient plus the solids effect.

3 2
 1   l  v ls
 
1/2 2/3 1/2
l g   C vt 
*
i m  i l  2  9 0 3 8 9  g  D p  R sd     
 v ls  2 g Dp

(6-45)
 
1/2 2/3
l g
*
l    R sd C vt
 il  90389  
1/2
g Dp 
1/2 v ls
  C vt 

The Darcy-Weisbach friction factor for a smooth pipe can be approached by:

1 2
 l     v ls  
 Dp  (6-46)

With:

 0 .0 8 9  0 .0 8 8
  0 .0 1 2 1 6 and  1   0 .1 5 3 7  D p   and  2   0 .2 0 1 3   v ls  (6-47)

For laboratory conditions both powers are close to -0.18, while for real life conditions with higher line speeds and
much larger pipe diameters this results in a power for the line speed of about α1=-0.155 and for the pipe diameter
of about α2=-0.175. This should be considered when analyzing the models for heterogeneous transport.

This gives for the Darcy-Weisbach friction factor in a dimensionless form:

 0 .1 5 5  0 .1 7 5  0 .1 5 5  0 .1 7 2
   
1/6
 l  0 .0 1 2 1 6   v ls   Dp  0 .1 2 3 3   v ls   g Dp l g  (6-48)

Substitution gives:

 
1/2 5/6
l g
*
 R sd C vt
1
im  il  11145    (6-49)
0 .6 7 2 0 .1 5 5
 
1/2 v ls
g  Dp   C vt  v ls

With the solids effect factor Sk (to compare with Fuhrboter)defined as:

C vt
im  il  S k  (6-50)
v ls

This gives for the solids effect factor Sk:

 
1/2 5/6
l g
*
 R sd
1
Sk  11145   (6-51)
0 .6 7 2 0 .1 5 5
 
1/2
g  Dp   C vt  v ls

Page 180 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Slurry Transport Models.

6.2.9. Nomenclature Early History & Empirical and Semi-Empirical Models.

a Correction factor Jufin Lopatin -


b Correction factor Jufin Lopatin -
CD Particle drag coefficient -
Cms Spatial concentration by mass -
Cv Volumetric concentration -
Cvs Volumetric spatial concentration -
Cvt Volumetric transport/delivered concentration -
Cx Inverse particle Froude number squared according to Durand & Condolios Frp-2 -
Cx,Gibert Inverse particle Froude number squared according to Gibert -
d Particle diameter m
d0 Average particle diameter Jufin Lopatin m
d10 Particle diameter at which 10% by weight is smaller m
d25 Particle diameter at which 25% by weight is smaller m
d50 Particle diameter at which 50% by weight is smaller m
d60 Particle diameter at which 60% by weight is smaller m
dm Mean particle diameter m
Dp Pipe diameter m
Dp,H Hydraulic diameter pipe cross section above bed m
Du Durand & Condolios constant (176-181) or (81-85) -
Erhg Relative excess hydraulic gradient -
fl Fanning friction factor liquid -
fm Fanning friction factor mixture
ELM Equivalent Liquid Model -
FL, FL,m Durand & Condolios Limit Deposit Velocity coefficient -
Frldv Flow Froude number at the Limit Deposit Velocity/critical velocity -
Frfl Flow Froude number -
Frp Particle Froude number 1/√Cx -
g Gravitational constant 9.81·m/s2
i Hydraulic gradient m.w.c./m
im Hydraulic gradient mixture m.w.c./m
iw,il Hydraulic gradient water/liquid -
K Durand & Condolios constant (176-181) or (81-85) -
K Constant others (Yagi, Babcock, etc.) -
K Wilson proportionality constant -
K Turian & Yuan constant -
K1 Newitt coefficient for heterogeneous transport (1100) -
K2 Newitt coefficient for sliding/moving bed (66) -
L, ΔL Length of the pipeline m
LDV Limit Deposit Velocity m/s
LSDV Limit of Stationary Deposit Velocity m/s
MHGV Minimum Hydraulic Gradient Velocity m/s
Ncr Zandi & Govatos parameter for Limit Deposit Velocity -
p Probability -
per Relative excess pressure -
Δp Head loss over a pipeline length ΔL kPa
Δpm Head loss of mixture over a pipeline length ΔL kPa
Δpl, Δpw Head loss of liquid/water over a pipeline length ΔL kPa
PSD Particle Size Diagram/Distribution -
Rsd Relative submerged density -
Sk Solids effect factor Fuhrboter spatial concentration m/s
Skt Solids effect factor Fuhrboter transport concentration m/s
u* Friction velocity m/s
vls Line speed m/s
vls,ldv Limit Deposit Velocity (often called critical velocity) m/s
vls,h-h Transition velocity heterogeneous vs. homogeneous according to Newitt m/s

Copyright © Dr.ir. S.A. Miedema TOC Page 181 of 304


Introduction Dredging Engineering.

vls,MHGV Minimum Hydraulic Gradient Velocity m/s


vmin Minimum gradient velocity m/s
vt Terminal settling velocity of particles m/s
vl Average liquid velocity m/s
vp Average velocity particle m/s
vs Average velocity solids m/s
x Abscissa -
y Ordinate -
α Power in Yagi equation -
α Darcy Weisbach friction factor constant -
α Power of concentration in Turian & Yuan equation -
α1 Darcy Weisbach friction factor power -
α2 Darcy Weisbach friction factor power -
β Power of Richardson & Zaki equation -
β Power of Fanning friction factor liquid in Turian & Yuan equation -
γ Power of drag coefficient in Turian & Yuan equation -
δ Power of Froude number in Turian & Yuan equation -
ρl Liquid density ton/m3
ρw Density of water ton/m3
ρm Mixture density ton/m3
λl Darcy-Weisbach friction factor liquid to wall -
μsf Friction coefficient for sliding bed (see also Srs) -
Φ Durand relative excess pressure as ordinate in different graphs -
ψ Durand abscissa, equations may differ due to historical development, later the -
relative submerged density has been added, sometimes the particle Froude number
is omitted
ψ Particle shape coefficient, usually near 0.7 -
ψ* Particle factor Jufin Lopatin -
νw,νl Kinematic viscosity of water/liquid m2/s
νm Kinematic viscosity of mixture with Thomas equation m2/s
νr Relative kinematic viscosity νm/νw -
ξ Particle shape factor -
ξ Slip ratio Yagi -

Page 182 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Slurry Transport Models.

6.3. The Wilson-GIW (1992) Model for Heterogeneous Transport.


6.3.1. The Full Model.
Assuming that 50% of the solids is moving in the bed by granular contact at a line speed of v50, and assuming a
friction coefficient μsf between the particles and the pipe wall, the friction force in a pipe with length ΔL is:

 sf
Fs f   A p  L  C v  s  l   g (6-52)
2

This gives an excess pressure due to the solids of:

Fsf  sf
pm  pl   L C v  s  l g (6-53)
A p 2

In terms of the hydraulic gradient this can be written as:

pm  pl  sf s  l   sf


im  il   C v   C v R sd (6-54)
l g  L 2 l 2

Wilson (1997) has defined a stratification ratio or relative solids effect, which tells which fraction of the particles,
is in suspension and which part is in the fixed or moving bed, supported by granular contact. Wilson (1997) gives
the following general equation for the head losses in hydraulic transport, where μsf equals the friction factor of a
sliding bed, which he has determined to be μsf=0.44. For the 50% case this gives:

M
im  il pm  pl  sf  v 50 
E rh g  R      (6-55)
R sd  C v  l  g   L  R sd  C v 2  v ls 

When the line speed vls equals the v50, the stratification ratio is 0.22 or half the friction coefficient μsf. This can be
written in terms of pressures instead of hydraulic gradient as:

M
 sf  v 50 
pm  pl  l  g  L    R sd  C v (6-56)
2  v ls 

This equation can be written in the more generic form, matching the notations of the other theories:

 2 M 
 sf  g  R sd  D p M  1 
pm  pl 1    v 50    Cv  (6-57)
 l  v ls  
 

For the line speed, where 50% of the particles is in granular contact, v50, Wilson gives the following equation:

8  60  d 
50
v 50  w 50   cosh   (6-58)
l  Dp 
 

When the power M equals 1, this equation has the same form as the equation of Durand & Condolios, Gibert,
Fuhrboter, Jufin Lopatin and Newitt et al. The power M depends on the grading of the sand and can be determined
by:

1/2

M  0 .2 5  1 3  
2
 (6-59)

Copyright © Dr.ir. S.A. Miedema TOC Page 183 of 304


Introduction Dredging Engineering.

The variance σ of the PSD (Particle Size Distribution), can be determined by some ratio between the v50 and the
v85:

  60  d 
 w  cosh 
85


85
 
 v 85  
Dp

  lo g    lo g   (6-60)
 v 50    60  d 
50
 w  cosh  
50
 Dp 
  

The terminal settling velocity related parameter w, the particle associated velocity, can be determined by:

1/3
w  0 .9  v t  2 .7   R s d  g   l  (6-61)

It seems this equation mixes the homogeneous and heterogeneous regimes. For very small particles the second
term gives a constant particle associated velocity, which matches homogeneous behavior at operational line speeds.
Since the homogeneous behavior does not depend on the particle size, this gives a constant or asymptotic particle
associated velocity.

6.3.2. The Simplified Wilson Model.


The model of Wilson can be simplified with some fit functions, according to:

0 .4 5  0 .2 5
0 .3 5  R sd    l ,a c t u a l 
v 5 0  3 .9 3   1 0 0 0  d 5 0      (6-62)
  
 1 .6 5   w ,20 

In which the particle diameter d50 is in m and the resulting v50 in m/s. The third term on the right had side is the
relative viscosity, the actual liquid viscosity divided by the viscosity of water at 20 degrees Centigrade. In normal
dredging practice this term is about unity and can be neglected.

The exponent M is given by the approximation:

1
  d 85  
M   ln    (6-63)

  d 50  

Figure 6-7 shows the Wilson et al. (1992) model in Erhg(il) coordinates. The black dotted line is the Erhg=μsf/2 line.
The yellow circle the point with vls=v50. The heterogeneous Wilson lines have powers of M=1.7, 1.0 & 0.25. So
heterogeneous lines always cross the yellow circle, the v50 point, and are rotated around this point depending on
the power M. The higher the power M, the steeper the line in this graph.

6.3.3. Generic Equation.


Based on equation (6-57) and equation (6-62) and the assumption M=1 for d85/d50=2.72 and assuming the liquid
is water of 20 degrees Centigrade and the solids are sand (quarts), an equation is derived to compare Wilson with
the other theories:

 2 M 
 sf  g  R sd  D p M  1 
im  il   1    v 50    Cv  (6-64)
 l  v ls  
 

For the line speed v50:

0 .3 5 0 .3 5
v 5 0  3 .9 3   1 0 0 0  d 5 0   4 4 .1   d 5 0  (6-65)

This gives:

Page 184 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Slurry Transport Models.

 2 M 
 sf  g  R sd  D p  1 
 
M
0 .3 5
im  il   1   4 4 .1   d 5 0    Cv  (6-66)
 l  v ls  
 

Substituting equation (6-65) in equation (6-66) gives:

 2 M 
 sf  g  R sd  D p 0 .3 5  M  1 
i m  i l   1  4 4 .1   d 50  Cv 
M
   (6-67)
 l  v ls  
 

With the friction coefficient of μsf=0.44, M=1 and some simplifications, this gives:

 g  R sd  D p
3 
0 .3 5  1 
im  i l   1  1 9 .4    d 50    Cv
 (6-68)
 l v
 ls  
 

Giving each term the dimension of velocity gives:

 g  R sd  D p
3 
0 .3 5 0 .1  1 
i m  i l   1  2 7 .6 4    g  d 50  l g    C v
 (6-69)
 l  v ls  
 

Or:

0 .3 5 0 .1 1
i m  i l  1 3 .8 2   g  d 5 0  l g  R sd C v (6-70)
v ls

The result is an equation where the excess pressure due to the solids is proportional to the pipe diameter Dp and
almost proportional to the cube root of the d50 of the sand. There is no direct relation with the terminal settling
velocity vt or the particle drag coefficient CD.

6.3.4. Conclusions & Discussion.


The basic form of the equation for heterogeneous transport, equation (6-55), is of the type; hydraulic gradient
mixture equals hydraulic gradient carrier liquid + solids effect. This implies that the hydraulic gradient of the
carrier liquid and the solids effect are independent like in the Fuhrboter (1961) model. This in contrary with the
Durand & Condolios (1952), Newitt et al. (1955) and the Jufin & Lopatin (1966) models. These are of the type;
hydraulic gradient mixture equals hydraulic gradient carrier liquid times 1 + solids effect. Equations (6-69) and
(6-70) show the Wilson equation for both types, with correct dimensions.
Basically the model shows straight lines in the Erhg(il) graph with the v50, 0.22 point as a pivot point. Depending
on the value of the power M, the straight line pivots around this point. Figure 6-7 shows these straight line for
powers of 0.25, 1.0 and 1.7. The power of 1.7 for uniform sands matches very well with the DHLLDV model for
uniform sands.
The model gives good results if the physics on which it is based occur, medium sized particles.
For very fine particles in a large pipe there will never be a sliding bed. The stationary bed will vaporize (erode)
with increasing line speed, probably without sheet flow, see Figure 6-6. The 50% stratification criterion is not valid
here because the sliding friction is not 100% mobilized resulting in a much lower friction.
Very large particles will, almost always, be more than 50% stratified, resulting in a sliding bed or sliding flow, so
the model is also invalid in this case, see Figure 6-8.

Copyright © Dr.ir. S.A. Miedema TOC Page 185 of 304


Introduction Dredging Engineering.

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il


1.000
Fixed Bed Cvs=c.

Sliding Bed Cvs=c.

Heterogeneous Flow
Relative excess hydraulic gradient Erhg (-)

Cvs=c.
Homogeneous Flow
Cvs=Cvt=c.
Resulting Erhg curve
Cvs=c.
Resulting Erhg curve
Cvt=c.
Limit Deposit Velocity
0.100
Ratio Potential/Kinetic
Energy
Wilson Erhg=0.22

Wilson Pivot Point

Wilson M=1.70 uniform

Wilson M=1.00 medium


graded
Wilson M=0.25 very
graded
Wilson SB modified
0.010
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)
© S.A.M. Dp=0.7620 m, d=0.20 mm, Rsd=1.59, Cv=0.300, μsf=0.440

Figure 6-6: The power in the Wilson et al. (1992) model, d=0.2 mm.

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il


1.000
Fixed Bed Cvs=c.

Sliding Bed Cvs=c.

Heterogeneous Flow
Relative excess hydraulic gradient Erhg (-)

Cvs=c.
Homogeneous Flow
Cvs=Cvt=c.
Resulting Erhg curve
Cvs=c.
Resulting Erhg curve
Cvt=c.
Limit Deposit Velocity
0.100
Ratio Potential/Kinetic
Energy
Wilson Erhg=0.22

Wilson Pivot Point

Wilson M=1.70 uniform

Wilson M=1.00 medium


graded
Wilson M=0.25 very
graded
Wilson SB modified
0.010
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)
© S.A.M. Dp=0.7620 m, d=2.00 mm, Rsd=1.59, Cv=0.300, μsf=0.440

Figure 6-7: The power in the Wilson et al. (1992) model, d=2 mm.

Page 186 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Slurry Transport Models.

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il


1.000
Fixed Bed Cvs=c.

Sliding Bed Cvs=c.

Heterogeneous Flow
Relative excess hydraulic gradient Erhg (-)

Cvs=c.
Homogeneous Flow
Cvs=Cvt=c.
Resulting Erhg curve
Cvs=c.
Resulting Erhg curve
Cvt=c.
Limit Deposit Velocity
0.100
Ratio Potential/Kinetic
Energy
Wilson Erhg=0.22

Wilson Pivot Point

Wilson M=1.70 uniform

Wilson M=1.00 medium


graded
Wilson M=0.25 very
graded
Wilson SB modified
0.010
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)
© S.A.M. Dp=0.7620 m, d=20.00 mm, Rsd=1.59, Cv=0.300, μsf=0.440

Figure 6-8: The power in the Wilson et al. (1992) model, d=20 mm.

In the 3 figures, the DHLLDV framework is used for comparison.

Wilson et al. (2006) suggests the following equation for full stratified flow, based on μsf=0.4 (or 0.44):

0 .2 5 0 .2 5
 4  v sm   v sm 
im  il  C vt R sd  2   sf   or im  il  C vt  R sd  2   sf  
 3  v ls   v ls  (6-71)
B a s e d o n  s f  0 .4 0 B a s e d o n  s f  0 .4 4

As is shown in Figure 6-8 there will be a discontinuity jumping from the heterogeneous model to the full stratified
model, even if the same power of 0.25 is applied. In fact if more than 50% of the solids is stratified, the v50 point
at Erhg=0.22 in the 3 graphs will never be reached.

Copyright © Dr.ir. S.A. Miedema TOC Page 187 of 304


Introduction Dredging Engineering.

6.3.5. Analysis of the v50 Equations.


The analysis in this chapter is based on the personal interpretation of the author of this book.

Figure 6-9 shows data of Durand & Condolios (1952) in the relative excess hydraulic gradient versus the line speed
graph. The advantage of this graph is that it show the solids effect in a double logarithmic graph, where curved
line become straight lines. The graph also contains the horizontal sliding bed curve (dark brown solid line), the
50% sliding bed curve (black dots), the ELM curve (dashed dark blue line) and the minimum v50 curve (light blue
solid line). The intersection line speed of the data series (extrapolated) with the 50% sliding bed curve, the v50,
increases with increasing particle diameter. The question is how does it increase?

Based on the assumption that Wilson et al. (1992) developed a model for operation conditions, meaning for
operational line speeds. Usually this means line speeds above the Limit Deposit Velocity, or slightly below. In the
case considered, see Figure 6-9, this gives line speeds of about 2.5 m/s up to 5.5 m/s. Now 3 cases can be
considered:
1. Fine particles, d<0.0002 mm.
2. Medium particles, 0.0002<d<1 mm.
3. Coarse particles, d>1 mm.
Fine particles are assumed to be transport according to the homogeneous flow regime under operational conditions.
So there is a minimum v50 for particles with a diameter smaller than d=0.0002 mm. The resulting minimum curve
is shown in Figure 6-9 (the solid light blue curve). This curve is close to the data of the d=0.0002 mm particles.
Since Wilson et al. (1992) used different data to calibrate their model, the match is not exactly. This minimum v50
can be determined by:

1/3 8
v 5 0 ,m in  2 .7   R s d  g   l   (6-72)
l

Relative excess hydraulic gradient Erhg vs. Line speed vls


1.000
Equivalent Liquid Model

Sliding Bed Cvs=c.


Relative excess hydraulic gradient Erhg (-)

Half Sliding Friction


Coefficient

d=0.20 mm, Cv=0.050

d=0.37 mm, Cv=0.050


0.100

d=0.89 mm, Cv=0.050

d=2.05 mm, Cv=0.050

d=4.20 mm, Cv=0.050

Minimum v50 Curve

0.010
1 10
Line speed vls (m/sec)

© S.A.M. Dp=0.1524 m, Rsd=1.585, Cvt=0.050, μsf=0.416

Figure 6-9: The data of Durand & Condolios (1952).

For medium sized particles, the v50 is increasing. Figure 6-9 shows this increase for particles with d=0.37 mm and
d=0.89 mm. Wilson et al. (1992) assumed this increase is proportional to the terminal settling velocity vt. This
gives:

Page 188 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Slurry Transport Models.

 0 .9  v 
1/3 8
v 50  t  2 .7   R s d  g   l  (6-73)
l

For coarse particles, the steepness of the decrease of the Erhg parameter with increasing line speed vls is less than
for medium particles. In fact it looks like the larger the particle, the less steep the decrease, where very coarse
particles could result in an almost horizontal curve. This also means that the intersection line speed of the
extrapolated data points with the 50% sliding bed curve is more far away. A horizontal curve has an intersection
v50 at infinity. Wilson et al. (1992) took this behavior into account by adding a term with the particle diameter to
pipe diameter ratio, according to:

 60  d 
 0 .9  v 
1/3 8
 2 .7   R s d  g   l  50
v 50   cosh   (6-74)
t
l  Dp 
 

The Darcy Weisbach friction coefficient λl indirectly gives the influence of the pipe diameter. The cosh term gives
the influence of the particle diameter to pipe diameter ratio. Figure 6-10 shows the resulting v50 curve for a
Dp=0.1524 m pipe (6 inch) (the thick black solid line).

The v50 versus the Particle diameter


8

Wilson Original
7

6
Wilson Simplified

5
Wilson Simplified
Modified
v50 m/s)

DHLLDV Simplified
3

2 DHLLDV Framework
Simplified

1
DHLLDV Framework

0
1.E-05 1.E-04 1.E-03 1.E-02
Particle diameter (m)
Dp=0.1524 m, Rsd=1.585, Cvs=0.110, μsf=0.416

Figure 6-10: The v50 curves for a Dp=0.1524 m pipe.

M
im  il pm  pl  sf v 50 
E rh g  R      (6-75)
R sd  C v  l  g   L  R sd  C v 2  v ls 

For uniform sands Wilson et al. (1992) found a power M=1.7 in the above equation, which works well for medium
sands, however for the coarser sands the effect of the cosh term in the v50 should be accompanied by a decrease of
this power as follows from the data in Figure 6-9. Durand & Condolios (1952) found that for coarse particles the
hydraulic gradient and thus the Erhg parameter does not really increase, based on the particle Froude number. The
definition of coarse is related to the pipe diameter by Wilson et al. (1992) being d=0.015·Dp or d=Dp/60. This
gives for the v50:

 0 .9  v 
1/3 8
v 5 0*  t*  2 .7   R s d  g   l   cosh 1  (6-76)
l

Copyright © Dr.ir. S.A. Miedema TOC Page 189 of 304


Introduction Dredging Engineering.

The line speed where the heterogeneous curve intersects with the sliding bed curve is about:

2
v ls *   v 50* (6-77)
3

Now assume this line speed is a constant, given a certain pipe diameter, the power M for particles of d>0.015·Dp
is:

ln  2 
M 
 3 v 50  (6-78)
ln   
 2 v 5 0* 

In these equations the * is used for the particle diameter d>0.015·Dp. This approach leads to a continuous behavior
of the heterogeneous flow regime.

The v50 versus the Particle diameter


8

Wilson Original
7

6
Wilson Simplified

5
Wilson Simplified
Modified
v50 m/s)

DHLLDV Simplified
3

2 DHLLDV Framework
Simplified

1
DHLLDV Framework

0
1.E-05 1.E-04 1.E-03 1.E-02
Particle diameter (m)
Dp=0.4000 m, Rsd=1.585, Cvs=0.110, μsf=0.416

Figure 6-11: The v50 curves for a Dp=0.4 m pipe.

Wilson et al. (1992) simplified the v50 equation to:

0 .3 5 0 .4 5  0 .2 5
 d 50   R sd    l ,a c tu a l 
v 50  3 .9 3        (6-79)
  
 0 .0 0 1   1 .6 5   w ,20 

Giving a v50 of 3.93 m/s for a 1 mm sand particle in 20º water. This simplified equation is also shown in Figure
6-10 (the solid red line). This equation does not contain the influence of the pipe diameter and only approaches the
original equation for medium sands. Later the simplified equation has been adjusted for particles with diameters
from 0.2 mm to 0.5 mm with a factor:

d h  0 .0 0 0 2
f  (6-80)
0 .0 0 0 5  0 .0 0 0 2

Page 190 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Slurry Transport Models.

This is also shown in Figure 6-10 as the Wilson simplified modified curve (dashed red line). Where the original
equation seems to overestimate the solids effect of fine particles, this modification seems to underestimate the
solids effect.
Figure 6-10, Figure 6-11 and Figure 6-12 show the original v50 method, the simplified and the simplified modified
equations. The effect of not having the pipe diameter influence in the simplified and the simplified modified
equations results in a deviation of these equations with the original method for large diameter pipes.

The v50 versus the Particle diameter


8

Wilson Original
7

6
Wilson Simplified

5
Wilson Simplified
Modified
v50 m/s)

DHLLDV Simplified
3

2 DHLLDV Framework
Simplified

1
DHLLDV Framework

0
1.E-05 1.E-04 1.E-03 1.E-02
Particle diameter (m)
Dp=1.2000 m, Rsd=1.585, Cvs=0.110, μsf=0.416

Figure 6-12: The v50 curves for a Dp=1.2 m pipe.

A better simplification of the v50 is achieved with the following equation (DHLLDV simplified, the solid green
line):

0 .4 5 0 .4 5  0 .2 5 0 .0 9 2
 d 50   R sd    l ,a c tu a l   Dp 
v 50  3          (6-81)
    0 .1 5 2 4 
 0 .0 0 0 5   1 .6 5   w ,20   

Figure 6-10, Figure 6-11 and Figure 6-12 show this equation as well. For a Dp=0.1524 m (6 inch) pipe, this
equation gives an almost identical v50 compared to the original method for particle diameters in the range of 0.1 to
0.8 mm. For the Dp=0.4 m pipe this range is larger, from 0.1 to 2 mm. For the Dp=1.2 m the range is even larger
than for the Dp=0.4 m pipe, the approximation is valid for particles with a diameter from 0.1 mm and above.

Still the overestimation of the v50 for small particles remains, as a result of the original assumption of Wilson that
there is a minimum v50. Figure 6-13 shows that also particles with a diameter smaller than 0.1 mm show
heterogeneous behavior with a v50 much smaller than this minimum v50. The correction with equation (6-80) seems
too drastic and not at the right location. The following correction factor in the range of particle diameters from 0.1
to 0.3 mm seems more appropriate and approaches the DHLLDV curves closely.

d h  0 .0 0 0 1
f  (6-82)
0 .0 0 0 3  0 .0 0 0 1

A better approach of the v50 however is the following equation, based on the DHLLDV Framework:

Copyright © Dr.ir. S.A. Miedema TOC Page 191 of 304


Introduction Dredging Engineering.

0 .0 9 2 1 .9 6 0 .4
 Dp   vt    l ,a c tu a l 
v 5 0  3 .6        (6-83)
 0 .1 5 2 4     
   g  d   w ,20 

Figure 6-10, Figure 6-11 and Figure 6-12 show the results of this equation and also the results of the much more
complicated DHLLDV Framework. In the range of particle diameters of 0.3 to 1 mm, this equation matches the
original Wilson method very closely. Above 1 mm the original Wilson method goes to infinity (because of the
decreasing power M), while the DHLLDV simplified equation goes to a constant value (assuming the power M
does not change) matching the findings of Durand & Condolios (1952). For particles with a diameter smaller than
0.3 mm this equation matches many experimental data and solves the problem of overestimation of the original
Wilson method in a more structural way.

6.3.6. Near Wall Lift.


Wilson et al. (2000) found that close to the bottom of the pipe the volumetric concentration is lower than just above
the bottom of the pipe. They explain this phenomena as the effect of near wall turbulent lift. If there is a strong
curvature of the velocity profile, which there is in turbulent flow close to the wall, particles in this flow will be
subject to a lift force. This should not be mistaken with Magnus lift in a laminar flow (in the viscous sub-layer).

Wilson et al. (2010) introduced the lift force FL on a particle as:

1 2  2
FL  C L   l  u*  d (6-84)
2 4

And for the weight of the particle FW:

s  l   g 
 3
FW  d (6-85)
6

Giving for the so called lift ratio LR, the ratio of the lift force on a particle FL to the weight of a particle FW:

2 2 2
FL 3
u* 3
l  u* 3 3
 l  v ls
LR   CL    CL   CL   CL   (6-86)
FW 4
R sd  g  d 4
 l  R sd  g  d 4 32
R sd  g  d

With: θ=Shields parameter.

This ratio will be referred to as the lift force to weight ratio or just the lift ratio. If this ratio is bigger than 1,
particles will be lifted, otherwise gravity dominates. Now it is the question whether or not the lift force is
completely correct, but it can be used as an indication. In Wilson et al. (2006) a lift coefficient CL=0.27 is
mentioned.

The line speed where the lift ratio equals 1 can now be determined with:

0 .5
2 32 R sd  g  d  3 2 R sd  g  d 
v ls , L 1    v ls , L 1     (6-87)
R
3 CL l R
 3 CL l 

Now the idea is that in some circumstances particles will be lifted away from the bottom of the pipe, leaving a
particle lean region near the wall, which is in agreement with the concentration profiles found. Medium sized
particles show a hydraulic gradient below the ELM curve after passing this curve, see Figure 6-13. The asymptotic
behavior for very high line speeds cannot be established from the data, since in general experiments are not carried
out with very high line speeds, but at line speeds close to operational line speeds or lower. Figure 6-14 shows data
from Blythe & Czarnotta (1995) also showing crossing the ELM. However these data points seem to return to the
ELM at higher line speeds, or at least follow a parallel curve, not approaching the pure liquid behavior.

Page 192 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Slurry Transport Models.

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il


Sliding Bed Cvs=c.
Mean

1.000 Equivalent Liquid Model


Relative excess hydraulic gradient Erhg (-)

Homogeneous Flow
Cvs=Cvt=c.

Resulting Erhg curve


Cvs=c.
0.100

Resulting Erhg curve


Cvt=c.

Limit Deposit Velocity

0.010 Ratio Potential/Kinetic


Energy

d=0.400 mm, Cv=0.170

d=0.085 mm, Cv=0.237

0.001
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)
© S.A.M. Dp=0.1016 m, d=0.085 mm, Rsd=1.585, Cv=0.237, μsf=0.416

Figure 6-13: The data of Whithlock et al. (2004).

Relative excess hydraulic gradient Erhg vs. Hydraulic gradient il


Sliding Bed Cvs=c.
Mean
Equivalent Liquid Model
1.000
Relative excess hydraulic gradient Erhg (-)

Homogeneous Flow
Cvs=Cvt=c.
Resulting Erhg curve
Cvs=c.
Resulting Erhg curve
Cvt=c.
0.100
Limit Deposit Velocity

Ratio Potential/Kinetic
Energy
Cv=0.285

Cv=0.235
0.010

Cv=0.185

Cv=0.135

Cv=0.085

0.001
0.001 0.010 0.100 1.000
Hydraulic gradient il (-)
© S.A.M. Dp=0.1000 m, d=0.280 mm, Rsd=1.585, Cv=0.175, μsf=0.416

Figure 6-14: The data of Blythe & Czarnotta (1995).

Based on the so called shear Reynolds number:

v*  d
R e*  (6-88)
l

Copyright © Dr.ir. S.A. Miedema TOC Page 193 of 304


Introduction Dredging Engineering.

A new expression has been derived for the stratification ratio R:

im  il 0 .9 3
E rh g   R  (6-89)
R sd  C vt    R e*
1/3

The coefficient is slightly different from the original paper (0.93 instead of 0.7) because here the Shields parameter
is applied. According to Wilson et al. (2010) the stratification ratio has an upper limit of 0.6. The shear velocity
used in this Reynolds number is the shear velocity based on the terminal settling velocity of the particle. Now
assuming a cylinder around the particle with diameter d and height d, the shear stress on the surface of this cylinder
follows from the weight of the particle and the surface of this cylinder, so:


s  l  g 
2 3
c    d  d
6
(6-90)
1
c  s  l   g  d
6

Assuming a similar relation between the shear stress and the shear velocity gives:

1
s  l   g 
2
c  l  v*  d
6

1/2
 s  l 
1/2
  R sd  g  d 
v*   g d    
 6 l 
   6  (6-91)

1/2
 R sd  g  d 
  d
v*  d  6 
R e*  
l l

Now substituting both the Shields parameter and the shear Reynolds number in the stratification ratio equation
gives:

im  il 0 .9 3
E rh g   R 
R sd  C vt  R 1/2 
1/3
 sd  g  d 
 d 
 2   
 l  v ls   6  
  
 8R   
 sd  g  d 
l
 
  (6-92)
 

1/2 1/3
1 0   R sd  g  d    R sd  g   l 
E rh g  R 
2
 l  v ls

With this equation, the proportionality with the line speed is about -1.8. The Shields parameter has a proportionality
of 2 because of the line speed squared and about -0.2 because of the Darcy Weisbach friction factor, resulting in a
power of -1.8. The shear Reynolds number has a power of 0. So this totals to -1.8, which gives a slightly steeper
decrease of the hydraulic gradient or stratification ratio of the original model having a power of -1.7 for uniform
PSD’s. The simplified heterogeneous model used:

M
im  il  sf  v 50 
E rh g  R     (6-93)
R sd  C v 2  v ls 

Page 194 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Slurry Transport Models.

With:

0 .4 5  0 .2 5
0 .3 5  R sd    l ,a c t u a l 
v 5 0  3 .9 3   1 0 0 0  d 5 0     
  
 1 .6 5   w ,20 
(6-94)
0 .3 5 0 .4 5  0 .2 5
v 5 0  1 .1 1   d 5 0  R sd  l 

For sands and gravels this reduces to:

0 .3 5
v 5 0  4 4 .1   d 5 0  (6-95)

Giving for the simplified equation (with μsf=0.44 and M=1.7):

M
 4 4 .1  d 0 .3 5 
 sf  50  d
0 .6
E rh g  R     1 3 7 .4  (6-96)
2  v ls  v ls
1 .7
 

The near wall lift based equation gives:

1/2
d
E rh g  R  (6-97)
2
 l  v ls

For medium sized sand particles in water and large pipe diameters (large flow Reynolds numbers), both equations
are close. For example, a 1 mm particle gives a v50 of 3.93 m/s, resulting in R=0.22 according to the simplified
method. The near wall lift method results in R=0.20, assuming λl=0.01 for large diameter pipes.
In both models the stratification ratio increases with increasing particle diameter and relative submerged density.
However the relation of the viscosity is different. The simplified model shows a decreasing stratification ratio with
increasing viscosity, while the near wal lift shows the opposite. For sands and gravels in water this will not have a
significant influence, but for other solids and liquids it might. The appearance of the Darcy Weisbach friction
factor in the denominator of the near wall lift equation results in some dependence of the pipe diameter. The
dependence on the sliding friction coefficient is not present anymore in the near wall lift equation. Using the
Shields parameter to explain for the stratification ratio seems interesting however.

Copyright © Dr.ir. S.A. Miedema TOC Page 195 of 304


Introduction Dredging Engineering.

6.3.7. The Demi-McDonald of Wilson (1979).


The LSDV of Wilson (1979) based on the 2 layer model was originally given as a nomographic chart, made with
the help of Professor F.M. Woods. This nomographic chart is known as the demi-McDonald, because of the shape
of the particle diameter curve. Figure 6-15 shows the demi-McDonald of Wilson (1979). The figure shows an
example of the LSDV for a Dp=0.25 m diameter pipe and a d=1 mm diameter particle. For solids with a density
of ρs=2.65 ton/m3 like sands and gravels, this gives an LSDV of about 3.1 m/s according to the left part of the
nomogram. The right part shows that solids with a density of ρs=7.85 ton/m3 like iron, give an LSDV of about 6.5
m/s.

Figure 6-15: The demi-McDonald of Wilson (1979).

The maximum Limit of Stationary Deposit Velocity (LSDV) vsm can be estimated by (Matousek (2004)), with d
in mm and Dp in m:

0 .5 5
  sf  R sd  0 .7 1 .7 5
8 .8    Dp d
 0 .6 6  v sm (6-98)
v sm  and FL 
2 0 .7
d  0 .1 1  D p
2  g  R sd  D p

Page 196 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Slurry Transport Models.

6.3.8. The Sliding Bed Regime New Developments.


Based on the developments of the DHLLDV Framework some new equations are derived for the sliding bed
regime.

First of all the relation between the spatial and the delivered concentration can be determined by, based on the slip
ratio:

 v ls   1 
C vs     C vt     C vt (6-99)
 v ls  v s l  1  

The slip ratio at the LSDV from the DHLLDV Framework can be estimated by the following empirical equation:

C vt
C vr 
C vb

(6-100)
   sf 2  0 .0 2 5 0 .6 5 
   0 .8 3    C v r  0 .5   0 .0 2 5  D p D p C v r 

 1  C vr   e    
4
 ls d v 

Now knowing the slip ratio, the spatial volumetric concentration can be determined at the LSDV (vsm) and with
this the approximation equation for the sliding bed hydraulic gradient according to:

0 .2 5 0 .2 5
C vs ( v sm )  v sm   1   v sm 
im  il   C vt  R sd   sf     il     C vt  R sd   sf    (6-101)
C vt  v ls   1     v ls 

This methodology would also be a good addition to the 4 component model, for the stratified fraction.

Copyright © Dr.ir. S.A. Miedema TOC Page 197 of 304


Introduction Dredging Engineering.

6.3.9. Nomenclature Wilson-GIW Models.

Ap Cross section pipe m2


A1 Cross section above bed m2
A2 Cross section bed m2
Cvb Volumetric spatial bed concentration -
Cvc Spatial volumetric concentration contact load -
Cvs Spatial volumetric concentration -
Cvs,1 Spatial volumetric concentration in cross section 1 -
Cvs,2 Spatial volumetric concentration in cross section 2 -
Cvs,f Spatial volumetric concentration homogeneous fraction -
Cvr Relative volumetric concentration Cvr=Cvs/Cvb -
Cvr,max Relative volumetric concentration at maximum LSDV -
Cvt Delivered (transport) volumetric concentration -
Cvt,f Delivered (transport) volumetric concentration homogeneous fraction -
Cvt,ph Delivered (transport) volumetric concentration pseudo homogeneous fraction -
Cvt,h Delivered (transport) volumetric concentration heterogeneous fraction -
Cvt,s Delivered (transport) volumetric concentration stratified fraction -
d Particle diameter m
d50 Particle diameter with 50% passing m
DH Hydraulic diameter m
Dp Pipe diameter m
Erhg Relative excess hydraulic gradient -
F Force kN
F1,l Force between liquid and pipe wall kN
F12,l Force between liquid and bed kN
F2,pr Force on bed due to pressure kN
F2,sf Force on bed due to friction kN
F2,l Force on bed due to pore liquid kN
Fn Normal force kN
Fw Weight of bed kN
Fsf Friction force, sliding kN
Fr Froude number -
g Gravitational constant 9.81 m/s2 m/s2
ibed Hydraulic gradient sliding bed m/m
ihom Hydraulic gradient homogeneous flow m/m
il Hydraulic gradient liquid m/m
im Hydraulic gradient mixture m/m
iplug Hydraulic gradient plug flow m/m
if Hydraulic gradient homogeneous fraction m/m
iph Hydraulic gradient pseudo homogeneous fraction m/m
ih Hydraulic gradient heterogeneous fraction m/m
is Hydraulic gradient stratified fraction m/m
ks Bed roughness m
ΔL Length of pipe section m
LDV Limit Deposit Velocity m/s
LSDV Limit of Stationary Deposit Velocity m/s
M Power stratification ratio between 0.25 and 1.7 -
n Porosity bed -
Op Circumference pipe m
O1 Circumference pipe above bed m
O2 Circumference pipe in bed m
O12 Width of bed m
Δp Pressure difference kPa
Δp1 Pressure difference on cross section 1 kPa
Δp2 Pressure difference on cross section 2 kPa
Δpl Pressure difference liquid kPa
Δpm Pressure difference mixture kPa

Page 198 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Slurry Transport Models.

Δpf Pressure difference homogeneous fluid kPa


q Power to determine the normalised excess hydraulic gradient -
Re Reynolds number -
Rsd Relative submerged density -
R Stratification ratio -
RH Hydraulic radius m
Sf Relative density homogeneous fraction + carrier liquid -
Sfp Relative density homogeneous & pseudo homogeneous fractions + carrier liquid -
Sfph Relative density homogeneous & pseudo homogeneous & heterogeneous fractions + -
carrier liquid
Sfphs Relative density homogeneous & pseudo homogeneous & heterogeneous & stratified -
fractions + carrier liquid
Sm Relative density mixture -
Ss Relative density solids -
u* Friction velocity m/s
v Velocity m/s
vs Limit of Stationary Deposit Velocity m/s
vt Terminal settling velocity m/s
vt* Dimensionless terminal settling velocity -
vu Threshold velocity m/s
vls Line speed m/s
v1, v1,m, vr Cross section averaged velocity above bed m/s
v2, v2,m, vb Cross section averaged velocity bed m/s
vsm Maximum Limit of Stationary Deposit Velocity (LSDV) m/s
v50 Line speed with 50% stratification m/s
v85 Line speed with 85% stratification m/s
vr Relative line speed vr=vls/vsm -
w50 Particle associated velocity matching the d50 m/s
w85 Particle associated velocity matching the d85 m/s
X Fraction in general -
Xf Homogeneous fraction -
Xph Pseudo homogeneous fraction -
Xh Heterogeneous fraction -
Xs Stratified fraction -
yb Height of bed m
α Multiplication factor bed friction -
α Power to determine LSDV -
β Power to determine LSDV -
β Bed angle rad
ε Pipe wall roughness m
ρl Density carrier liquid ton/m3
ρs Density solids ton/m3
ρm Mixture density ton/m3
ρx Density mixture with fraction X ton/m3
ρf Density homogeneous fluid ton/m3
ρfp Density homogeneous+pseudo homogeneous fluid ton/m3
ρfph Density homogeneous+pseudo homogeneous+heterogeneous fluid ton/m3
ρfphs Density homogeneous+pseudo homogeneous+heterogeneous+stratified fluid ton/m3
θ Shields parameter -
θc Critical Shields parameter -
λ Darcy-Weisbach friction factor -
λl Darcy-Weisbach friction factor liquid-pipe wall -
λ1 Darcy-Weisbach friction factor with pipe wall -
λ2 Darcy-Weisbach friction factor with pipe wall, liquid in bed -
λ12 Darcy-Weisbach friction factor on the bed -
λf Darcy-Weisbach friction factor based on homogeneous fluid properties -
νl Kinematic viscosity m2/s
νl,actual Actual kinematic viscosity liquid m2/s

Copyright © Dr.ir. S.A. Miedema TOC Page 199 of 304


Introduction Dredging Engineering.

νw,20 Kinematic viscosity of water at 20 degrees centigrade m2/s


νf Kinematic viscosity homogeneous fluid m2/s
τ Shear stress kPa
τl Shear stress liquid-pipe wall kPa
τ1,l Shear stress liquid-pipe wall above bed kPa
τ12,l Shear stress bed-liquid kPa
τ2,l Shear stress liquid-pipe in bed kPa
τ2,sf Shear stress from sliding friction kPa
μsf Sliding friction coefficient -
μl Dynamic viscosity liquid Pa·s
μf Dynamic viscosity homogeneous fluid Pa·s
μr Relative dynamic viscosity -
σn Normal stress kPa
ζ Normalised excess hydraulic gradient -
ζ∞ Normalised excess hydraulic gradient at infinite line speed -

Page 200 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Slurry Transport Models.

6.4. Notes.

Copyright © Dr.ir. S.A. Miedema TOC Page 201 of 304


Introduction Dredging Engineering.

Page 202 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Slurry Transport Models.

Copyright © Dr.ir. S.A. Miedema TOC Page 203 of 304


Introduction Dredging Engineering.

Page 204 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Slurry Transport Models.

Copyright © Dr.ir. S.A. Miedema TOC Page 205 of 304


Introduction Dredging Engineering.

Page 206 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Modeling of the Swing Winches of a Cutter Dredge.

Chapter 7: Modeling of the Swing Winches of a Cutter Dredge.

7.1. Introduction.
The dredge motions consist of the six degrees of freedom of the pontoon complemented with the rotation of the
ladder around the ladder bearings. This gives a total of 7 degrees of freedom (surge, sway, heave, roll, pitch, yaw
and ladder rotation). For a dredge operating in still water, when wave forces are ignored, the motions in the
horizontal plane are relevant (surge, sway and yaw) as well as the ladder rotation. The three pontoon motions can
be reduced to the rotation around the spud if the spud is considered to be infinitely stiff. If the ladder rotation is
considered not to be the result of a mass-spring system, but controlled by the ladder winch, only one equilibrium
equation has to be solved, the rotation of the pontoon around the spud. The other 6 equilibrium equations are of
interest when working offshore, when wave forces have to be taken into account, but using these equations
increases the calculations to be carried out enormous.

7.2. The Motions of the Dredge.


The equilibrium equation of rotation around the spud is a second order non-linear differential equation, with the
following external forces:

 The inertial forces of pontoon and ladder


 The water damping on pontoon and ladder
 The spring forces resulting from the swing wires
 The external forces resulting from the current
 The external forces resulting from the cutting process
 The external forces resulting from the swing winches
 The external forces resulting from the pipeline
 The reaction forces on the spud

Figure 7-1: The display of the top view of the cutterdredge, also showing the channel.

Copyright © Dr.ir. S.A. Miedema TOC Page 207 of 304


Introduction Dredging Engineering.

1. The inertial forces (moments) determine whether there is an acceleration or deceleration of the rotation around
the spud. These forces are the result of the equilibrium equation and thus of the external forces.
2. The water damping and the current forces depend on the value and the direction of the current and on the
rotational speed of the pontoon around the spud.

3. The spring forces resulting from the swing wires and the forces resulting from the swing winches strongly
depend on the characteristics of the winches and the wires and the winch control system. The position of the
anchors in relation to the position of the spud and the position of the swing wire sheaves on the ladder
determines the direction of the swing wire forces and thus of the resulting moments around the spud. Figure
4 shows the winch output of a research simulator.

Figure 7-2: The display of the back view of the cutterdredge,


also showing the cross-sectional channel profile.

4. The forces and moment excerted on the pontoon by the current influence the rotation around the spud
depending on the current speed and the swing speed. For small values of the current speed this effect can
however be neglected. For high values of the current speed the influence depends on the direction of the
current and the swing angle. It may occur that the swing winches do not have enough power to pull back the
pontoon out of a corner due to the angle of the swing wires and a high current speed.

5. The cutting forces and the cutting torque strongly influence the rotation around the spud, these will be
discussed in the paragraph concerning the cutting forces.
6. The winch forces and the winch moment strongly influence the rotation around the spud, these will be
discussed in the paragraph concerning the swing winch characteristics.

7. The forces resulting from the pipeline can be neglected if the position of the swivel elbow is close to the
position of the work spud, because in this case this force hardly influences the rotation of the pontoon around
the spud.

8. The reaction forces on the spud can be determined by the equilibrium equations of forces and complement
this equilibrium. These forces however do not contribute to the moment around the spud.

The rotation of the pontoon around the spud is dominated by the cutting forces, the winch characteristics, the
inertia of pontoon and ladder and placement of the anchors, while damping and current play a less important role.
The equilibrium equation can be formulated as:

I yaw   s  k yaw   s  c yaw   s  M cu rren t  M c u ttin g  M w ir e s  M p ip e  M sp u d (7-1)

The water damping is combined with the current moment, the wire spring force, the pipeline moment and the spud
moment are not taken into consideration. Equation 1 thus reduces to:

I yaw   s  M cu rren t  M c u ttin g  M w ir e s (7-2)

The equilibrium equation in question is non-linear, while some of the data is produced by interpolation from tables.
This implies that the equation will have to be solved in the time domain, using a certain time step. This is also
necessary because the simulation program has to interact with the console (the user input). To simulate the motions

Page 208 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Modeling of the Swing Winches of a Cutter Dredge.

of the dredge real time, a time step of at least two times per second is required. A time step of 5 to 10 times per
second would be preferred.

Figure 7-3: The display of the side view of the cutterdredge,


also showing the longitudinal channel profile.

7.3. The Influence of the Swing Angle on the Wire Moment.


With fixed anchor positions, the angle of attack of the swing wires relative to the axis system of the pontoon,
changes continuously with the value of the swing angle. With large swing angles this may result in a large decrease
of the effective pulling or braking moment of the swing wires. This decrease of course depends on the anchor
positions relative to the pontoon.

In this paper the following coordinate system definitions are applied:

1. The origin is placed in the centerline of the work spud.


2. The two wire sheaves are positioned on the centerline through the work spud and the cutterhead.
3. The positive swing direction is counter clock wise, with an angle of zero degrees when the centerline of the
dredge matches the vertical axis (y-axis).
4. The distance from the center of the workspud to the center of the sheaves is Lss.

With the coordinates if the swing sheaves on the ladder xss and yss according to:

x s s  L s s  s in   s  (7-3)

And

y ss  L ss  c o s   s  (7-4)

The length of the port wire and the angle of the port wire with the centerline of the channel can be determined
according to:

y 
2 2
L pw   x ss  x p w   ss
 y pw (7-5)

and

 x  x 
ss pw
 p w  a r c ta n   (7-6)
 y  y 
 ss pw 

Copyright © Dr.ir. S.A. Miedema TOC Page 209 of 304


Introduction Dredging Engineering.

S e s s io n s : P o rt w in c h ro p e fo rc e v s tim e
C o m p a n y: S A M - C o n s u lt
N a m e
G ra p h 4
1 00 0

8 00

6 00

k N 4 00
: D r. ir . S .A . M ie d e m a
: W in c h e s T im e R e c o r d in g s
1

2 00

0
0 0:00 0 0:01 0 0:02 0 0:03 0 0:04 0 0:05 0 0:06 0 0:07 0 0:08 0 0:09 0 0:10
T im e

P o rt w in c h ro p e s p e e d v s tim e
6 0.0

3 6.0
m /m in

1 2.0

-1 2 . 0

-3 6 . 0

-6 0 . 0
0 0:00 0 0:01 0 0:02 0 0:03 0 0:04 0 0:05 0 0:06 0 0:07 0 0:08 0 0:09 0 0:10
T im e

S ta rb o a rd w in c h ro p e s p e e d v s tim e
C o p y R ig h t: S A M - C o n s u lt
D re d g e : C u t te r S u c t io n D re d g e
M a y 2 6 , 1 9 9 8 , 1 0 :3 2 :0 8 P M
C u tt e r S u c t io n D r e d g e S im u la t o r V 3 . 0 2

6 0.0

3 6.0
m /m in

1 2.0

-1 2 . 0

-3 6 . 0

-6 0 . 0
0 0:00 0 0:01 0 0:02 0 0:03 0 0:04 0 0:05 0 0:06 0 0:07 0 0:08 0 0:09 0 0:10
T im e

S ta rb o a rd w in c h ro p e f o rc e v s tim e
1 00 0

8 00

6 00
k N

4 00

2 00

0
0 0:00 0 0:01 0 0:02 0 0:03 0 0:04 0 0:05 0 0:06 0 0:07 0 0:08 0 0:09 0 0:10
T im e

Figure 7-4: The output of the winch parameters.

Figure 7-5: The coordinate system with the dredge in the neutral position.

Page 210 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Modeling of the Swing Winches of a Cutter Dredge.

The length of the starboard wire and the angle of the starboard wire with the centerline of the channel can be
determined according to:

 y ss  y sw 
2 2
L sw   x ss  x sw   (7-7)

and

 x ss  x sw 
 s w  a r c ta n   (7-8)
 y  y 
 ss sw 

Figure 7-6: The coordinate system with the dredge at a swing angle s.

The angle of the port wire with the centerline of the dredge is:

 pw   s (7-9)

The angle of the starboard wire with the centerline of the dredge is:

 sw   s (7-10)

The moment around the spud, resulting from the forces in the swing wires can now be determined by:

M w ir e s  
 F p w  L s s  s in  p w   s  F s w  L s s  s in   s w   s  (7-11)

The relation between the rope speed of the port wire and the angular speed of the dredge is now:


 s  L s s  v p w  s in  p w   s  (7-12)

The relation between the rope speed of the starboard wire and the angular speed of the dredge is now:

Copyright © Dr.ir. S.A. Miedema TOC Page 211 of 304


Introduction Dredging Engineering.

 s  L s s  v s w  s in   s w   s  (7-13)

This results in loss of effective power of both winches. The power mobilized by the winches to the angular speed
of the dredge is:

2
 
2
P wm  P p w m  P s w m  F p w  v p w  s in  p w   s  F s w  v s w  s in   s w   s  (7-14)

The power consumed by the winches is:

P w  P p w  P sw  F p w  v p w  F sw  v sw (7-15)

7.4. The Winch Characteristics.


The torque speed characteristic of the winches consists of two parts if an electric drive is assumed. The first part
runs from 0 revolution up to full revolutions and has a linear decrease of the torque, from a maximum at zero
revolutions to the full torque at full revolutions. At this last point also the full power of the drive is reached. At
higher revolutions the drive will use field weakening, while the power stays constant. In the simulator it is assumed
that the characteristics for hauling and braking are equal. If one winch is in hauling mode, the other one will always
be in braking mode.

Figure 7-7: The torque-speed characteristic of the winches.

7.5. The Control System of the Winches.


The hauling winch is controlled by a setpoint for the winch revolutions. The braking winch is controlled by a
setpoint for the braking torque. So for the hauling winch, the available torque results from the revolutions, while
the pulling force also results from the drumdiameter and the number of layers on the drum. The mobilized torque
also depends on the loads (cutter and current) and on the angular acceleration of the dredge around the spud pole.

Figure 7-8 shows the actual revolutions of the hauling winch, the setpoint of the hauling winch, the setpoint of the
braking winch and the load curve for the hauling winch. The load curve includes the cutting process, the current
and water damping and the braking winch. The difference between the available torque and the torque resulting
for the loads is available for the acceleration of the pontoon. In the example given in Figure 7-8, it is assumed that
the actual revolutions of the winch are smaller then the setpoint and that the available torque is larger then the
required torque for compensating the loads.

The actual torque mobilized by the hauling winch, is always the resulting torque necessary to reach or stay on the
setpoint. If in a certain situation, the torque available is less then the torque required, then the available maximum
torque is assumed. In this case the working point is the intersection point of the load curve with the vertical dotted
line through the setpoint of revolutions. The maximum available torque is not fully mobilized.

Page 212 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Modeling of the Swing Winches of a Cutter Dredge.

Figure 7-8: The torque-speed characteristic of the winches with the setpoints.
Case where the required torque is sufficient.

Figure 7-9: The torque-speed characteristic of the winches with the setpoints.
Case where the required torque in the setpoint is not sufficient.

Figure 7-10: The torque-speed characteristic of the winches with the setpoints.
Case where the setpoint is smaller then the actual revolutions.

Copyright © Dr.ir. S.A. Miedema TOC Page 213 of 304


Introduction Dredging Engineering.

Figure 7-11: The dredge, winch and channel layout.

Figure 7-12: The dredge and anchor layout Figure 7-13: The dredge and anchor layout
for case 1, port. for case 1, starboard.

Figure 7-9 shows the case where the winch torque required in the setpoint is not sufficient. In this case, the working
point is the intersection point of the load curve with the torque-speed curve. The maximum available torque is
fully mobilized. The setpoint is not reached because there is not sufficient torque available.
Figure 7-10 shows the case where the setpoint is smaller then the actual revolutions. In this case, the pontoon will
decelerate. The working point is the intersection point of the vertical through the setpoint and a minimum torque
required keeping the wire from going slack.

Page 214 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Modeling of the Swing Winches of a Cutter Dredge.

7.6. Case Studies.


To show the behavior of the dredge-winch system two cases will be shown. In the first case the dredge starts on
the centerline of the channel. The dredge and winch layouts are shown in Figure 7-11.

7.6.1. Case 1:
The winches have a drum diameter of 0.84 m, a full power of 158 kW at 8.87 rpm. The resulting full torque is 167
kNm. The anchor positions are symmetrical with respect to the centerline and are 65 m in horizontal direction and
-21.5 m in vertical direction, away from the sheaves on the ladder. The ladder is not in contact with the bank and
is moving free through the water. See Figure 7-12 and Figure 7-13.

The following actions are taken:

1. The setpoint for the swingspeed is set to 24 m/min to starboard.


2. The dredge swings from 0 to 30 degrees to starboard.
3. The setpoint for the swingspeed is set to 24 m/min to port.
4. The dredge swings from 30 degrees starboard to 30 degrees port.
5. The setpoint for the swingspeed is set to 24 m/min to starboard.
6. The dredge swings from 30 degrees port to the centerline.

Figure 7-14 shows the rope speeds and pulling forces for both the port and the starboard winch. It is clearly shown
in the graphs in Figure 7-14 that, while the rope forces increase instantly, the rope speed increases or decreases
according to a first or second order system. This is caused by the mass-spring-damper system according to equation
1, but also by the inertia of the winches themselves. In the simulator, the winches are modeled as a first order
system. The winches and the dredge need some time to accelerate or decelerate.

The deceleration requires more time in case 1 then the acceleration, because the braking force is set to 30% of the
maximum force, which is about 180 kN. The pulling force however, can be much higher, depending on the
characteristic of the winches. Setting the braking force to a higher value, will increase the speed of the deceleration.

Typical for this case is, that the pulling wire is more and more perpendicular to the ladder when the swing angle
approaches 30 degrees. This results in a decreasing pulling force, which can be seen in Figure 7-14. The braking
force is set to a constant value and will only differ from this value if the braking force is larger then the torque-
speed curve permits it to be. In that case the braking force will follow the torque speed curve.

7.6.2. Case 2:
The winches have a drum diameter of 0.84 m, a full power of 158 kW at 8.87 rpm. The resulting full torque is 167
kNm. The anchor positions are symmetrical with respect to the centerline and are 65 m in horizontal direction and
+3.5 m in vertical direction, away from the sheaves on the ladder, as is shown in Figure 7-16 and Figure 7-17.
The ladder is not in contact with the bank and is moving free through the water.

The following actions are taken:

1. The setpoint for the swingspeed is set to 24 m/min to starboard.


2. The dredge swings from 0 to 30 degrees to starboard.
3. The setpoint for the swingspeed is set to 24 m/min to port.
4. The dredge swings from 30 degrees starboard to 30 degrees port.
5. The setpoint for the swingspeed is set to 24 m/min to starboard.
6. The dredge swings from 30 degrees port to the centerline.

Figure 7-15 shows the rope speeds and pulling forces for both the port and the starboard winch. Because the
anchors are moved 25 m forward in the channel, now the angle between the pulling wire and the ladder decreases
when the dredge approaches the 30 degrees swing angle. This results in an increase of the pulling force as is visible
in Figure 7-15. The start and stop behavior is almost equal to case 1.

Copyright © Dr.ir. S.A. Miedema TOC Page 215 of 304


Introduction Dredging Engineering.

Figure 7-14: The rope speeds and forces for case 1.

Figure 7-15: The rope speeds and forces for case 2.

Page 216 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Modeling of the Swing Winches of a Cutter Dredge.

Figure 7-16: The dredge and anchor layout Figure 7-17: The dredge and anchor layout
for case 2, port. for case 2, starboard.

7.7. Conclusions.
The modeling of the winches and the wires consists of solving the equilibrium equation of motions of the dredge
around the spudpole in combination with the characteristics of the winches. The two cases show that it takes about
10 seconds to accelerate to a swing speed of 24 m/min. The time required for the deceleration is of the same
magnitude, but depends of course on the setpoint of the brake force.

The two cases also show, that the shape rope speed and force as a function of time, strongly depend on the position
of the anchors relative to the sheave positions at the ladder. The two cases describe symmetrical configurations,
which of course is not always the case. An infinite number of configurations can be chosen. Which configuration
is the best depends on the work to be carried out and on the boundary conditions of the work to be carried out.

Copyright © Dr.ir. S.A. Miedema TOC Page 217 of 304


Introduction Dredging Engineering.

7.8. Nomenclature.
cyaw Spring constant of the yaw motion kNm/rad
Fpw Rope force of the port wire kN
Fsw Rope force of the starboard wire kN
Iyaw Mass moment of inertia of pontoon in yaw direction kNms2/rad
kyaw Damping coefficient of pontoon in yaw direction kNms/rad
Lpw Length of the port wire m
Lss Distance from working spud to swing sheaves on ladder m
Lsw Length of starboard wire m
Mcurrent Moment around the spud exerted by the current kNm
Mcutting Moment around the spud exerted by the cutting process kNm
Mpipe Moment around the spud exerted by the floating pipeline kNm
Mspud Moment around the spud exerted by the spud kNm
Mwires Moment around the spud exerted by the swing wires kNm
nfull Full revolutions of the swing winch rpm
Ppw Power of the port winch kW
Ppwm Power of the port winch mobilized on the dredge kW
Psw Power of the starboard winch kW
Pswm Power of the starboard winch mobilized on the dredge kW
Pw Power of both winches kW
Pwm Power of both winches mobilized on the dredge kW
Tacc Winch torque available for acceleration or deceleration kNm
Tfull Full torque of the winches kNm
Tmax Maximum torque of the winches kNm
vpw Rope speed of the port winch m/sec
vsw Rope speed of the starboard winch m/sec
xpw X coordinate of the port anchor m
xss X coordinate of the swing sheaves on the ladder m
xsw X coordinate of the starboard anchor m
ypw Y coordinate of the port anchor m
yss Y coordinate of the swing sheaves on the ladder m
ysw Y coordinate of the starboard anchor m
s Swing angle rad
pw Port wire angle rad
sw Starboard wire angle rad

Page 218 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Modeling of the Swing Winches of a Cutter Dredge.

7.9. Notes.

Copyright © Dr.ir. S.A. Miedema TOC Page 219 of 304


Introduction Dredging Engineering.

Page 220 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Modeling of the Swing Winches of a Cutter Dredge.

Copyright © Dr.ir. S.A. Miedema TOC Page 221 of 304


Introduction Dredging Engineering.

Page 222 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Modeling of the Swing Winches of a Cutter Dredge.

Copyright © Dr.ir. S.A. Miedema TOC Page 223 of 304


Introduction Dredging Engineering.

Page 224 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

Chapter 8: The Trailing Suction Hopper Dredge.

8.1. Introduction
In the last decennia there has been a strong development in the enlargement of TSHD’s (Trailing Suction Hopper
Dredges) from roughly 10.000 m3 in the early 90’s up to 50.000 m3 expected loading capacity nowadays. Because
of the economy of the loading process, but also environmental regulations, it is important to predict the overflow
losses that are occurring.
For the estimation of the sedimentation process in TSHD’s a number of models have been developed. The oldest
model used is the Camp (1936), (1946) and (1953) model which was developed for sewage and water treatment
tanks. Camp and Dobbins (1944) added the influence of turbulence based on the two-dimensional advection-
diffusion equation, resulting in rather complicated equations. Miedema (1981) used the Camp model to develop
an analytical model. Groot (1981) added the effects of hindered settling. Vlasblom & Miedema (1995) and
Miedema & Vlasblom (1996) simplified the Camp equations by means of regression and included a rising sediment
zone, as well as hindered settling and erosion and an adjustable overflow. Van Rhee (2002C) modified the
implementation of erosion in the Camp model, but concluded that the influence is small due to the characteristics
of the model. Ooijens (1999) added the time effect, since the previous models assume an instantaneous response
of the settling efficiency on the inflow of mixture. Yagi (1970) developed a new model based on the concentration
distribution in open channel flow.
The models mentioned above are all black box approaches assuming simplified velocity distributions and an ideal
basin. Van Rhee (2002C) developed a more sophisticated model, the 2DV model. This model is based on the 2D
(horizontal and vertical) Reynolds Averaged Navier Stokes equations with a k-ε turbulence model and includes
suspended sediment transport for multiple fractions.

8.2. The Loading Cycle of a Hopper Dredge


The loading cycle of a TSHD is considered to start when the hopper is filled with soil and starts to sail to the dump
area. This point in the loading cycle was chosen as the starting point in order to be able to show the optimal load
in a graph. The loading cycle then consists of the following phases:
 Phase 1: The water above the overflow level flows away through the overflow. The overflow is lowered to
the sediment level, so the water above the sediment can also flow away. In this way minimum draught is
achieved. Sailing to the dump area is started.

Figure 8-1: Phase 1 of the loading cycle.

 Phase 2: Continue sailing to the dump area.

Figure 8-2: Phase 2 of the loading cycle.

Copyright © Dr.ir. S.A. Miedema TOC Page 225 of 304


Introduction Dredging Engineering.

 Phase 3: Dump the load in the dump area. Dumping can be carried out in 3 different ways, using the bottom
dumping system, pumping ashore or rain bowing.

Figure 8-3: Phase 3 of the loading cycle.

 Phase 4: Pump the remaining water out of the hopper and sail to the dredging area. Often the water is not
pumped out, but instead water is pumped in, to have the pumps as low as possible, in order to dredge a higher
density, which should result in a shorter loading time.

Figure 8-4: Phase 4 of the loading cycle.

 Phase 5: Start dredging and fill the hopper with mixture to the overflow level, during this phase 100% of the
soil is assumed to settle in the hopper.

Figure 8-5: Phase 5 of the loading cycle.

 Phase 6: Continue loading with minimum overflow losses, during this phase a percentage of the grains will
settle in the hopper. The percentage depends on the grain size distribution of the sand.

Figure 8-6: Phase 6 of the loading cycle.

Page 226 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

 Phase 7: The maximum draught (CTS, Constant Tonnage System) is reached. From this point on the
overflow is lowered.

Figure 8-7: Phase 7 of the loading cycle.

 Phase 8: The sediment in the hopper is rising due to sedimentation, the flow velocity above the sediment
increases, resulting in scour. This is the cause of rapidly increasing overflow losses.

Figure 8-8: Phase 8 of the loading cycle.

Figure 8-9 and Figure 8-10 show the total load, the effective load, the TDS and the overflow losses during these
phases. The way each phase occurs in the cycle, depends on the type of hopper dredge, the working method and
of course, the type of soil to be dredged.

Basically there are two main methods for loading the hopper. The ‘Constant Volume System’ (CVS). This system
has a fixed overflow level so the effective volume of the hopper is constant. The TSHD is designed for filling the
hopper with sediment with a density of 1.9-2.0 ton/m3. The ‘Constant Tonnage System’ (CTS). The system has an
adjustable overflow level. The hopper is designed for a density of 1.3-1.7 ton/m3 in combination with a maximum
tonnage. When the content of the hopper reaches the maximum tonnage, the overflow is lowered in order to keep
the tonnage of the hopper content constant. This system has certain advantages, like reaching the maximum
tonnage sooner than with CVS, resulting in the pumps to be as low as possible, giving a higher mixture density.
De Koning (1977) has compared both systems.
The sedimentation in the hopper occurs during the phases 5, 6, 7 and 8. During phase 5 the hopper is filled with
mixture until the overflow level is reached. During this phase 100% of the soil is assumed to stay in the hopper
and settle. When the overflow level is reached, phase 6, depending on the grain distribution, a specified percentage
of the soil will not settle and will leave the hopper via the overflow. During this phase scouring does not have
much influence on the sedimentation process. When the maximum weight of the hopper contents is reached, the
overflow will be lowered continuously in order to keep the weight of the hopper contents constant at its maximum
(only CTS system). When the sediment level rises, phase 8, the flow velocity above the sediment increases and
scouring will re suspend settled particles. The overflow losses increase with time. The transition between phase 5
and 6 is very sharp, as is the transition between the phases 6 and 7 for the graph of the total load, but this does not
exist in the graph of the effective load (Figure 8-10). However, the transition between the phases 7 and 8 is not
necessarily very sharp. When this transition occurs depends on the grain distribution of the soil dredged. With very
fine sands this transition will be near the transition between phases 6 and 7, so phase 7 is very short or may not
occur at all. With very coarse sands and gravel scouring is minimal, so phase 8 is hardly present. In this case the
sediment level may be higher than the overflow level. With silt the phases 7 and 8 will not occur, since after
reaching the overflow level the overflow losses will be 100%.

Copyright © Dr.ir. S.A. Miedema TOC Page 227 of 304


Introduction Dredging Engineering.

T he hopper dredge cycle.


5000
phase 5
phase 6
4500
phase 7
phase 8
phase 1
4000
phase 2
phase 3
3500

3000
L o a d in t o n s

2500

2000

1500

1000

500

phase 4
0
-300 -275 -250 -225 -200 -175 -150 -125 -100 -75 -50 -25 0 25 50 75 100
T ime in min

Total load Effective (situ) load Tonnes D ry Solids Overflow losses M axim um production

Figure 8-9: The loading cycle of a TSHD.

T he loading curves for an 0.3 mm d50 sand.


5000

4500

phase 7 phase 8
4000

phase 6
3500

3000
L o a d in t o n s

2500

phase 5
2000

1500

1000

500

0
0.0 4.2 8.4 12.6 16.8 21.0 25.2 29.4 33.6 37.8 42.0
T ime in min

Total load Effective (situ) load Tonnes D ry Solids Overflow losses M axim um production

Figure 8-10: The loading part of the cycle of a TSHD.

Page 228 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

So far the total load in the hopper has been described. A contractor is, of course, interested in the "Tonnes Dry
Solids" (TDS) or situ cubic meters. The total load or gross load consists of the sediment with water in the pores
and a layer of water or mixture above the sediment. The TDS consists of the weight of the soil grains only. The
net weight in the hopper consists of the weight of the sediment, including the weight of the pore water. If the
porosity of the sediment is considered to be equal to the in-situ porosity, then the volume of the sediment in the
hopper equals the removed situ-volume. Although, in practice, there will be a difference between the in-situ
porosity and the sediment porosity, here they will be considered equal. The net weight (weight of the sediment
Ws) is equal to the weight in the hopper Wh minus the weight of the water above the sediment Ww:

Ws = Wh - Ww (8-1)

The net volume (volume of the sediment Vs) is equal to the volume of the hopper Vh minus the volume of the
water above the sediment Vw.

Vs = Vh - V w (8-2)

Multiplying the volumes with the densities gives:

Vs   s = W h - V w   w and V w = Vh - Vs (8-3)

V s   s = W h - (V h - V s )   w (8-4)

V s  (  s -  w )= W h - V h   w (8-5)

Rearranging the terms of equation (8-5) gives an expression for the volume of situ cubic meters.

(W h - V h   w )
Vs = (8-6)
( s -  w )

Multiplying the situ volume Vs with the situ density ρs gives for the situ weight Ws:

(W h - Vh   w )   s
W s = Vs  s = (8-7)
( s -  w )

To find the weight of the sand grains only (without the pore water), the situ density ρs has to be replaced by the
quarts density (or particle density) ρq:

s  w q (W h - V h   w )   q
TD S=Ws   = (8-8)
q  w s ( q -  w )

The net weight (situ weight) according to equation (8-7) can be approximated by the total weight of the load in the
hopper minus the weight of the same volume of water and the result multiplied by 2. For the TDS this factor is
about 1.2, according to equation (8-8). This is of course only valid for a specific density of the sediment of 2 tons
per cubic meter.

With these equations the hopper cycle for the net weight and the TDS can be derived, this is shown in Figure 8-9
and Figure 8-10. The hopper dredge is optimally loaded, when the effective load (weight) or the TDS divided by
the total cycle time dWs/dt reaches its maximum. This is shown in Figure 8-9 and Figure 8-10 and is the reason
for the starting point of the loading cycle in Figure 8-9.

Copyright © Dr.ir. S.A. Miedema TOC Page 229 of 304


Introduction Dredging Engineering.

8.3. The Calculation Model.


Consider a rectangular hopper of width W, height H and length L. A mixture with a mixture density ρm and with
a specified grain distribution is being dredged. Depending on the operational conditions such as dredging depth,
the pump system installed, the grain distribution (PSD, Particle Size Distribution) and mixture density ρm, a
mixture flow Q will enter the hopper. If the porosity n of the sediment is known, the flow of sediment can be
determined according to:

The mass flow of the mixture into the hopper is:

Q in   m = Q in  (  w  (1 -C v ) +  q  C v ) (8-9)

The mass flow of the solids into the hopper is now:

dTDS ( m -  w )
 Q in q  = Q in  C v q (8-10)
dt ( q -  w )

From this, the mass flow of situ sediment into the hopper is:

dWs
= Q in  C v  (  q + e   w ) (8-11)
dt

With:

n
e  (8-12)
(1  n )

Part of this mass flow will settle in the hopper and another part will leave the hopper through the overflow. The
ratio between these parts depends on the phase of the loading process. During phase 5 the hopper is loaded to the
overflow level, so the mass flow into the hopper will stay in the hopper. This means that the total settling efficiency
ηb during this phase equals 1. During phase 6 the loading continues until the maximum load in the hopper is
reached (CTS). If scouring does not occur, the mass flow that will settle into the sediment can be calculated with
equation (8-13) and (8-14), where the settling efficiency ηb should be determined with equation (8-56) and (8-57),
0.

The mass flow of the solids staying in the hopper is now:

dTDS
 Q in  C v   q   b (8-13)
dt

From this, the mass flow of situ sediment into the hopper is:

dWs
= Q in  C v  (  q + e   w )   b (8-14)
dt

During phase 7 the loading continues, but with a CTS, the overflow is lowered to ensure that the total weight in
the hopper remains constant. As scour does not yet occur, the above equation is still valid. During phase 8 scouring
occurs. If scouring does occur, the mass flow that will settle into the sediment can also be calculated with equation
(8-13) and (8-14), but the settling efficiency should be determined with equation (8-56) and (8-57) taking into
account the effect of scouring. Scouring is the cause of increasing overflow losses. Scour depends upon the velocity
of the flow above the sediment. Since in a hopper the sediment is not removed, the sediment level rises during the
loading of the hopper. This means that the height of the mixture flow above the sediment decreases during the
loading process, resulting in an increasing flow velocity. The scour velocity can now be determined by:

Q in
ss  (8-15)
B H w

Page 230 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

The height of the water/mixture layer Hw above the sediment, is equal to the overflow height H minus the sediment
height Hs:

Ws
H w  H  Hs  H  (8-16)
s W L

The overflow height H is a constant for a Constant Volume System (CVS), but this height changes for a CTS,
because the overflow is lowered from the moment, the maximum weight in the hopper is reached. If a maximum
weight Wm is considered, the height of the layer of water above the sediment Hw for a CTS can be determined by:

Wm    H s  B  L
s
H w  (8-17)
w B L

The hopper loading curve can now be determined by first calculating the time required to fill the hopper (phase 6),
given a specified mixture flow Qin. From the mixture density m the mass and given a specified porosity, the
volume of the sediment can be calculated. From this point the calculations are carried out in small time steps
(phases 7 and 8). In one time step, first the height of the sediment and the height of the water layer above the
sediment are determined. The height of the water layer can be determined with equation (8-16) for a CVS hopper
and equation (8-17) for a CTS hopper. With equation (8-15) the scour velocity can now be determined. Using
equations (8-55) the fraction of the grains that will be subject to scour can be determined. If this fraction ps is zero
equation (8-50) has to be used to determine the mass flow that will stay in the hopper. If this fraction is not equal
to zero equation (8-56) has to be used. Equations (8-13) and (8-14) can now be used to determine the mass flow.
This mass flow multiplied by the time step results in an increment of the sediment mass that is added to the already
existing mass of the sediment. The total sediment mass is the starting point for the next time step. This is repeated
until the overflow losses are 100%. When the entire loading curve is known, the optimum loading time can be
determined. This is shown in Figure 8-9, where the dotted line just touches the loading curve of the effective (situ)
load or the TDS. The point determined in this way gives the maximum ratio of effective load or TDS in the hopper
and total cycle time. In chapter 2 and chapter 3 the determination of the settling efficiency ηb will be discussed in
detail.

8.4. The Layer Thickness of the Layer of Water above Overflow Level
Where an obstacle is constructed on the bottom of an open channel, the water surface is raised and passes over it.
Structures of this type are called weirs. Aside from special cases, flow over weirs may be regarded as steady, i.e.
unchanging with respect to time, and suddenly varied, as in most hydraulic structures. The most important problem
arising in connection with weirs is the relationship between the discharge over the weir and the characteristics of
the weir. Many authors have suggested various relationships (e.g. Poleni, Weissbach, Boussinesq, Lauck, Pikalow)
generally along the same theoretical lines and with similar results. So it seems satisfactory to introduce only the
relationship of Weissbach.

 
3 2 3 2
v   v 
2 2
2
Q out  Ce b  2g h       (8-18)
3  2g   2g  
 

If h/(M+h) tends towards zero (because h is small compared to M) then v2/2gh also tends towards zero; so a
simplified relationship can be reached as introduced first by Poleni about 250 years ago:

2
Q out  Ce b h  2g h (8-19)
3

The above equation (8-19) gives the relation between the layer thickness h and the flow Qout
for the stationary process. During the dredging process of a TSHD however, the process is not
always stationary. At the start of the loading process when the overflow level is reached the
layer of water will build up, while at the end when the pumps stop the layer thickness will
decrease to zero. If the TSHD makes turns and the poor mixture is pumped overboard directly,

Copyright © Dr.ir. S.A. Miedema TOC Page 231 of 304


Introduction Dredging Engineering.

also the layer thickness will decrease and as soon as the mixture is pumped back in the hopper
the layer will build up again.

Figure 8-11: A sharp crested weir.

Figure 8-12: Values for the coefficient Ce as a function of ha/hb=h/M.

First the increase of the layer thickness will be considered. This increase per unit of time multiplied by the width
and the length of the hopper equals the difference between the flow into the hopper and the flow leaving the hopper
through the overflow according to:

dh
bL   Q in  Q o u t (8-20)
dt

Substituting equation (8-19) in this equation gives a non-linear differential equation of the first order for the layer
thickness h.

Page 232 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

dh 2
bL   Q in  C e   2g b h
3/2
(8-21)
dt 3

This equation can be solved numerically, for example in Excel, using the starting condition t=0, h=0 and the
following two equations:

2
Q in  C e   2g b h
3/2

3 (8-22)
h   t
b L
h i1  h i   h (8-23)

5000

4500

4000

3500

3000

2500 Load
Volume
2000 TDS

1500

1000

500

0
12:00:00
12:14:05
12:28:10
12:42:15
12:56:20
13:10:25
13:24:30
13:38:35
13:52:40
14:06:45
14:20:50
14:34:55
14:49:00
15:03:05
15:17:10
15:31:15
15:45:20
15:59:25
16:13:30
16:27:35
16:41:40
16:55:45
17:09:50
17:23:55
17:38:00
17:52:05
18:06:10
18:20:15
18:34:20
18:48:25
19:02:30
19:16:35
19:30:40
19:44:45
19:58:50
-500

Figure 8-13: An example of a loading cycle of a TSHD with many turns.

In the equilibrium situation where Qin=Qout, the maximum layer thickness hmax is found according to:

2/3
  2/3
 Q in   Q in 
h m ax       (8-24)
 C 
2
  2 .9 5  C  b 
 e 2g b 
e

 3 

From the start, t=0, until the maximum layer thickness is reached, hmax, the layer thickness h is a function of time
that can be approximated according to:

t
 1/3

 
2/3
  2 .9 5  C e  b 
0 .4 5 2  L     
t

Q in  
1  e   h
 Q 
h (t)   
in
m ax
1 e   (8-25)
 2 .9 5  C e
 b     
 
 

1/3 (8-26)
 2 .9 5  C e  b  1/2
  0 .4 5 2  L     0 .4 5 2  L  h m ax
 Q in 

The decrease of the layer thickness h when the pumps are stopped or the poor mixture is pumped overboard follows
from equation (8-20) when Qin is set to zero, this can be approximated by:

Copyright © Dr.ir. S.A. Miedema TOC Page 233 of 304


Introduction Dredging Engineering.

2 3/2 (8-27)
Ce   2g b h
3
h   t
b L
h i1  h i   h (8-28)

Solving this gives:

h m ax ( 3 .2 7  0 .0 4 8 6  b )  1 .2 8 4
h ( t )  h m ax  with: Cd  L (8-29)
1  C 
0 .2 2
 h m ax  t
2/3 4/3
d
b

Figure 8-15 shows the discharge and the loading of the layer of water above the overflow level for a hopper with
a length of 40 m, a width of 9 m and a height of 9 m and a flow of 5.8 m 3/sec. Both the exact solution and the
approximation are shown versus an in situ measurement. The effective width of the overflow is assumed to be
equal to the width of the hopper.

Figure 8-14: A close up of the hopper volume registration.

Loading & discharge of layer of water


0.6

0.5
Layer thickness (m)......

0.4

Loading
Loading approximation
0.3
Discharge
Discharge approximation
Measurement
0.2

0.1

0
0 100 200 300 400 500 600

Time (sec)

Figure 8-15: The layer thickness during a turn, registration and approximation.

Page 234 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

M e a s u re d v s c a lc u la te d lo a d in g c y c le
5000

T o ta l lo a d (M )

V o lu m e (M )
4500 T D S (M )

T o ta l in (M )

O v e rflo w T D S (M )
4000 T o ta l lo a d (C )

V o lu m e (C )

T D S (C )
L o a d (to n ) , V o lu m e (m 3 )

3500 T o ta l in (C )

O v e rflo w T D S (C )

T D S b u ffe re d

3000 O v e rflo w T D S b u f fe re d

2500

2000

1500

1000

500

0
0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150 160 170 180 190

T im e (m in )

Figure 8-16: The cycle as registered is simulated with the theoretical model.

T h e lo a d in g cu rve s.
9000 9000

8100 8100

7200 7200

6300 6300

5400 5400

L o a d in to n s
L o a d in to n s

4500 4500

3600 3600

2700 2700

1800 1800

900 900

0 0
0 5 10 15 20 25 30 35 40 45 50 55 60 65 70

T ime in min

Total load Effective (situ) load Tonnes D ry Solids Overflow losses effective Overflow losses TD S

Figure 8-17: The decreasing of the height of the layer of water above the overflow at the end of the cycle.

Copyright © Dr.ir. S.A. Miedema TOC Page 235 of 304


Introduction Dredging Engineering.

8.5. The Storage Effect.


In the Miedema & Vlasblom model (1996) upon entrance of a particle in the hopper it is decided whether the
particle will settle or not. In reality the particles that will not settle first have to move through the hopper before
they reach the overflow. This means that these particles are part of the TDS in the hopper during the time they stay
in the hopper. Ooijens (1999) discovered that using the time delay to determine the overflow losses improved the
outcome of the Miedema & Vlasblom model (1996) considerably. Overflow losses with time delay can be derived
from the overflow losses without a time delay according to the following equation:

t t 
1 s
ov b (t) 

  ov c (t)  dt 

  ( o v c ( t )  o v b ( t ))  d t (8-30)
t  0

The first term in equation (8-30) gives the time delay for the situation with a constant bed height. Since the height
of the bed increases during the loading process, the rising bed pushes part of the mixture out of the hopper. This is
represented by the second term on the right hand.

T h e lo a d in g c u rv e s o f th e s m a l l T S H D .
4 00 0 4 00 0

3 60 0 3 60 0

3 20 0 3 20 0

2 80 0 2 80 0

2 40 0 2 40 0

L o a d in t o n s
L o a d in t o n s

2 00 0 2 00 0

1 60 0 1 60 0

1 20 0 1 20 0

8 00 8 00

4 00 4 00

0 0
0 .0 2 .5 5 .0 7 .5 1 0.0 1 2.5 1 5.0 1 7.5 2 0.0 2 2.5 2 5.0 2 7.5 3 0.0 3 2.5 3 5.0 3 7.5 4 0.0 4 2.5

T im e in m i n
Load TDS Miedema Overflow TDS Miedem Load TDS van Rhee Overflow TDS van Rhe TDS in Miedema TDS in van Rhee Overflow TDS buffer Load TDS buffer Sediment TDS

Figure 8-18: Loading curves according to Miedema & van Rhee (2007) with and without time delay.

Figure 8-18 shows the loading and overflow curves with and without the time delay or storage effect for a case
considered by Miedema & van Rhee (2007). Table 8-1 gives the main data of the TSHD used in this case.

Table 8-1: The data of the TSHD used.


Hopper Load Volume Length Width Empty Flow Hopper Mixture
height load v0 density
ton m3 m m m m3/sec m/sec ton/m3
Small 4400 2316 44.0 11.5 4.577 4 0.0079 1.3

From top to bottom Figure 8-18 contains 9 curves. The first two curves (blue and green) are almost identical and
represent the TDS that enters the hopper. Since the flow and the density are constant, these curves are straight. The
3rd curve (red) represents the total TDS in the hopper according to the Miedema & Vlasblom (1996) model, so
including the TDS that is still in suspension above the sediment of which part will leave the hopper through the
overflow. The 4th curve (green) represents this according to van Miedema & van Rhee (2007). The 5th curve (blue)

Page 236 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

represents the TDS that will stay in the hopper excluding the time delay effect, according to Miedema & Vlasblom
(1996). The 6th (brown) curve represents the TDS in the sediment in the hopper. The 7th curve (blue) is the overflow
losses according to Miedema & Vlasblom (1996), so excluding the time delay or buffering effect. The 8 th curve
(green) represents the overflow losses according to the 2DV model of van Rhee (2002C), which automatically
includes the time delay effect. The 9th curve (red) represents the overflow losses according to the Miedema &
Vlasblom (1996) model including the time delay effects according to equation (8-30).

8.6. The Hopper of a TSHD as an Ideal Settlement Basin.


As stated before, the ideal settlement basin is a rectangular basin with an entrance zone, a settlement and
sedimentation zone and an overflow zone. The hopper geometry and configuration aboard of the TSHD can be
quite different from the ideal situation, so a method to schematize the hopper dimensions is required.

1. The height H of the hopper can be defined best as the hopper volume divided by the hopper area LW. This
means that the base of the ideal hopper, related to the maximum overflow height is at a higher level than the
ship's base. This assumption results in a good approximation at the final phases (7 and 8) of the loading
process, while in phase 6 of the loading process the hopper is filled with mixture and so the material stays in
the hopper anyway.

2. Near the loading chute of the hopper or in cases where a deep loading system is used, the turbulence of the
flow results in a good and sufficient distribution of the concentration and particle size distribution over the
cross-section of the hopper, so the entrance zone can be kept small. For example between the hopper bulkhead
and the end of the loading chute.

3. In the ideal settlement basin there are no vertical flow velocities except those resulting from turbulence.
However in reality vertical velocities do occur near the overflow, therefore it is assumed that the overflow
zone starts where the vertical velocities exceed the horizontal velocities. An estimate of where this will occur
can easily be made with a flow net.

4. Although the presence of beams and cylinder rods for the hopper doors does increase the turbulence, it is the
author’s opinion, that an additional allowance is not required, neither for the hopper load parameter, nor for
the turbulence parameter.

5. As is shown in Figure 8-6 and Figure 8-7, a density current may occur during the loading phases 6 and 7,
resulting in a non-uniform velocity and density distribution. This does not affect the so called hopper load
parameter as is proven in 0, so for the schematization of the hopper a uniform velocity and density distribution
are assumed.

6. The validity of the schematizations and simplifications will be proven by some examples with model and
prototype tests.

Copyright © Dr.ir. S.A. Miedema TOC Page 237 of 304


Introduction Dredging Engineering.

Page 238 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

8.7. The Modified Camp Model.


Sedimentation is a treatment process where suspended particles, like sand and clay are re-moved from the water.
Sedimentation can take place naturally in reservoirs or in compact settling installations. Sedimentation is applied
in groundwater treatment installations for backwash water treatment and in TSHD’s. In horizontal flow settling
tanks water is uniformly distributed over the cross-sectional area of the tank in the inlet zone. A stable, non-
turbulent, flow in the settling zone takes care for the settling of suspended matter in the settling zone. The sludge
accumulates on the bottom, or is continuously removed. In the outlet zone the settled sludge must be prevented
from being re-suspended and washed out with the effluent. Sedimentation occurs because of the difference in
density between suspended particles and water. The following factors influence the sedimentation process: density
and size of suspended particles, water temperature, turbulence, stability of flow, bottom scour and flocculation:
 Density, the higher the density of the particles, the faster the particles settle
 Size, the larger the particles are, the faster they settle
 Temperature, the lower the temperature of the water is, the higher the viscosity is, so the slower the
particles settle
 Turbulence, the more turbulent the flow is, the slower the particles settle
 Stability, instability can result in short circuit flow, influencing the settling of particles
 Bottom scour, by bottom scour settled particles are re-suspended and washed out with the effluent

Figure 8-19: The top view of the ideal basin.

Figure 8-20: The side view of the ideal basin.

The ideal settlement basin consists of an entrance zone where the solid/fluid mixture enters the basin and where
the grain distribution is uniform over the cross-section of the basin, a settlement zone where the grains settle into
a sediment zone and a zone where the cleared water leaves the basin, the overflow zone. It is assumed that the
grains are distributed uniformly and are extracted from the flow when the sediment zone is reached. Each particle
stays in the basin for a fixed time and moves from the position at the entrance zone, where it enters the basin
towards the sediment zone, following a straight line. The slope of this line depends on the settling velocity v and
the flow velocity above the sediment so. Figure 8-19 shows a top view of the ideal settlement basin. Figure 8-20
shows the side view and Figure 8-21, Figure 8-22 and Figure 8-23 the path of individual grains. All particles with
a diameter do and a settling velocity vo will settle, a particle with this diameter, entering the basin at the top, reaches
the end of the sediment zone. Particles with a larger diameter will all settle, particles with a smaller diameter will
partially settle. Miedema & Vlasblom (1996) adapted the Camp model to be used for hopper sedimentation. The
biggest difference between the original Camp (1936), (1946) and (1953) model and the Miedema & Vlasblom

Copyright © Dr.ir. S.A. Miedema TOC Page 239 of 304


Introduction Dredging Engineering.

model is the height Hw above the sediment zone. In the Camp model this is a fixed height, in the Miedema &
Vlasblom model this height decreases during the loading process.

Figure 8-21: The path of a particle with a settling velocity greater than the hopper load parameter.

Figure 8-22: The path of a particle with a settling velocity equal to the hopper load parameter.

Figure 8-23: The path of a particle with a settling velocity smaller than the hopper load parameter.

The average horizontal velocity so in the basin, when the height Hw above the sediment is known (see equations
(8-16) and (8-17)), equals to:

Q in
so  (8-31)
W H w

The hopper load parameter vo is defined as the settling velocity of a particle that enters the basin (hopper) at the
top and reaches the sediment at the end of the basin, after traveling a distance L, see Figure 8-22. This can be
determined according to (with a uniform velocity distribution):

vo H w H w Q in
 thus: v o  so   (8-32)
so L L W L

If the velocity distribution is non-uniform, like in Figure 8-24, the hopper load parameter can be derived by
integrating the horizontal velocity s(z) over the time the particle, entering at the top of the basin, needs to reach
the sediment at the end, so traveling a horizontal distance L.

Page 240 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

 s(z )  d t  L (8-33)
0

With:

H w
H w
T  , z  v 'o  t , d z  v 'o  d t , Q in  W   s(z )  d z (8-34)
v 'o 0

Equation (8-33) can be written as:

H w
1 1 Q in
  s(z )  d z  
W
 L (8-35)
v 'o 0
v 'o

Thus the hopper load parameter does not change because of a non-uniform velocity distribution.

Q in
v 'o   vo (8-36)
W L

During the transport of a particle from the top of the inlet to the overflow however, the sediment level rises by
ΔH=vsed·Δt, where Δt equals the traveling time of the particle and vsed equals the sediment (bed) rise velocity. The
thickness of the layer of fluid above the sediment thus decreases from Hw when the particle enters the hopper to
Hw-ΔH when the particle reaches the sediment at the end of the hopper due to the settling velocity of the particle.
The average thickness Ha of the layer of water above the sediment during the transport of the particle is now:

H a  H w  0 .5   H (8-37)

Figure 8-24: The path of a particle with a non-uniform velocity distribution.

The average horizontal velocity so in the hopper during the stay of the particle in the hopper is thus:

Q in Q in
so   (8-38)
W  (H w  0 .5   H ) W H a

The time it takes for the particle to be transported over the length of the hopper is thus:

L W L H a
t   (8-39)
so Q in

The vertical distance traveled by a particle that enters the hopper at the top and just reaches the sediment at the end
of the hopper is (see Figure 8-25):

W L H a
v oo   t  v oo   H w  H  H a  0 .5   H (8-40)
Q in

Copyright © Dr.ir. S.A. Miedema TOC Page 241 of 304


Introduction Dredging Engineering.

This gives for the settling velocity of such a particle:

Q in Q in  0 .5   H 
v oo   ( H a  0 .5   H )  1   (8-41)
W L Ha W L  Ha 

With:

W L H a
 H  v sed   t  v sed  (8-42)
Q in

This gives for the modified hopper load parameter:

Q in v sed
v oo   (8-43)
W L 2

A smaller hopper load parameter means that smaller grains will settle easier. From Figure 8-21 the conclusion can
be drawn that grains with a settling velocity greater than vo will all reach the sediment layer and thus have a settling
efficiency ηg of 1. Grains with a settling velocity smaller then vo, Figure 8-23, will only settle in the sedimentation
zone, if they enter the basin below a specified level. This gives for the modified settling efficiency of the individual
grain:

 vs 
 gg    (8-44)
 v oo 

Figure 8-25: The effect of a rising sediment level.

In the case of a non-uniform velocity distribution, Figure 8-24, the settling efficiency can also be defined as the
ratio of the horizontal distances traveled in the time a particle needs to reach the sediment, although this is not
100% true because the ratio of the vertical distance traveled gives the exact settling efficiency, it's a good
approximation:

 Lv 
g  
o
 (8-45)
 L 
 v 

The horizontal distance traveled by a particle in the time to reach the sediment level is:

Lv   s( z )  d t (8-46)
0

With:

Page 242 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

H w
H w
T 
vs
, z  vs t , dz  vs dt , Q in  W   s(z )  d z (8-47)
0

Equation (8-47) can be written as:

H w
1 1 Q in

vs
  s(z )  d z 
vs

W
 L v (8-48)
0

This also gives a settling efficiency according to:

 vs 
g    (8-49)
 vo 

The settling efficiency of a particle with a settling velocity smaller than the hopper load parameter vo, does not
change due to a non-uniform velocity distribution. If the fraction of grains with a settling velocity greater than vo
equals po, then the settling efficiency for a grain distribution ηb can be determined by integrating the grain settling
efficiency for the whole grain distribution curve, according to Figure 8-26. The blue surface equals the basin
settling efficiency according to equation (8-50).

Figure 8-26: Determination of the basin settling Figure 8-27: A graphical method to determine
efficiency. the settling efficiency.

po

b  1  p o   g  dp (8-50)
0

In theory a particle is removed from the water when it reaches the bottom of the settling tank. In practice, however,
it is possible that re-suspension of already settled particles occurs.
When the sediment level in the hopper is rising, the horizontal velocity increases and there will be a point where
grains of a certain diameter will not settle anymore due to scour. First the small grains will not settle or erode and
when the level increases more, also the bigger grains will stop settling, resulting in a smaller settling efficiency.
The effect of scour is taken into account by integrating with the lower boundary ps. The fraction ps is the fraction
of the grains smaller then ds, matching a horizontal velocity in the hopper of ss.
The shear force of water on a spherical particle is:

Copyright © Dr.ir. S.A. Miedema TOC Page 243 of 304


Introduction Dredging Engineering.

1 1 2
      w  ss (8-51)
4 2

The shear force of particles at the bottom (mechanical friction) is proportional to the submerged weight of the
sludge layer, per unit of bed surface (see Figure 8-28):

f    N    (1  n )  (  q   w )  g  d (8-52)

In equilibrium the hydraulic shear equals the mechanical shear and the critical scour velocity can be calculated.
The scour velocity for a specific grain with diameter ds, according to Huisman (1973-1995) and (1980) is:

8    (1  n )  (  q -  w )  g  d s
ss = (8-53)
 w

Figure 8-28: The equilibrium of forces on a particle.

With μ·(1-n)=0.05 and λ=0.03 this gives:

4 0  ( q -  w )  g  d s
ss = (8-54)
3 w

The particle diameter of particles that will not settle due to scour (and all particles with a smaller diameter) is:

3 w 2
ds =  ss (8-55)
4 0  ( q -  w )  g

Knowing the diameter ds, the fraction ps that will not settle due to scour can be found if the PSD of the sand is
known. Equation (8-54) is often used for designing settling basins for drinking water. In such basins scour should
be avoided, resulting in an equation with a safety margin. For the prediction of the erosion during the final phase
of the settling process in TSHD’s a more accurate prediction of the scour velocity is required, which will be
discussed in another chapter. The settling efficiency ηg, but this only occurs at the end of the loading cycle, can
now be corrected for scour according to:

po

b  1  po    g  dp (8-56)
ps

When ps>po this results in:

b  1  p s (8-57)

Page 244 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

8.8. The Influence of Turbulence.


For the ideal settlement basin laminar flow is assumed. Turbulent flow will reduce the settling velocity of the
grains and thus the total settling efficiency. Whether turbulent flow occurs, depends on the Reynolds number of
the flow in the basin. Using the hydraulic radius concept this number is:

Q in
Re  (8-58)
  (W  2  H w )

For a given flow Qin and viscosity  the Reynolds number depends on the width W and the height Hw of the layer
of fluid in the basin. A large width and height give a low Reynolds number. However this does not give an attractive
shape for the basin from an economical point of view, which explains why the flow will be turbulent in existing
basins.

Dobbins (1944) and Camp (1946) and (1953) use the two-dimensional turbulent diffusion equation to determine
the resulting decrease of the settling efficiency.

2
c  c    z  c 2
 c
s(z )   z   v (c )      x  (8-59)
x z
2
 z  z x
2

Assuming a parabolic velocity distribution instead of the logarithmic distribution, neglecting diffusion in the x-
direction and considering the settling velocity independent of the concentration reduces the equation to:

2
c  c c
s t  k  (h  z )
2
  x  z
2
 v
z
(8-60)
z

Because of the parabolic velocity distribution, the turbulent diffusion coefficient εz is a constant. A further
simplification is obtained if the velocity s is assumed constant throughout the depth, meaning that the constant of
the parabola k approaches zero. In this case the turbulent diffusion equation becomes:

2
c c  c c
 s  z  v (8-61)
t x z
2 z

Huisman (1973-1995) in his lecture notes derives the diffusion-dispersion equation in a more general form,
including longitudinal dispersion.

c  (s  c )   c    c 
  x   v c  z  (8-62)
t x x  x  z   z 

Assuming a steady and uniform flow, the longitudinal dispersion coefficient is independent of x and the settling
velocity v independent of z. This reduces the equation 18 to:

c c
2 2
 c  c
s  z   v  x  (8-63)
x z
2 2
z x

By means of computations Huisman (1973-1995) shows that the retarding effect of dispersion may be ignored for
the commonly applied width to depth ratio 3 to 5. This reduces equation (8-62) to equation (8-59) of Dobbins and
Camp.

Groot (1981) investigated the influence of hindered settling and the influence of different velocity distributions
using the following equation:

c c  v (c ) c   c 
s  v (c )   c     (x, z )   (8-64)
x z c z z  z 

Copyright © Dr.ir. S.A. Miedema TOC Page 245 of 304


Introduction Dredging Engineering.

The velocity distribution, the diffusion coefficient distribution and the distribution of the initial concentration did
not have a significant influence on the computed results, but the results were very sensitive on the formulation of
hindered settling. This formulation of course influences the settling velocity in general. Equation (8-63) can be
solved analytically using separation of variables. The boundary conditions used by Camp and Dobbins describe
the rate of vertical transport across the water surface and the sediment for x= and the concentration distribution
at the inlet, these are:

c
  v c  0 at the water surface (8-65)
z
c
  v c  0 at the sediment for x=, for the no-scour situation (8-66)
z
c  f z at the entrance for x=0 (8-67)

This method, resulting in Figure 8-29, Figure 8-30 and Figure 8-31, gives the removal ration due to turbulence for
a single grain. The removal ratio can be determined by summation of a series.
Solving equation (8-64) gives (vH/2z) as the independent parameter on the horizontal axis and the removal ratio
(v/vo=settling efficiency) on the vertical axis. Using a parabolic velocity distribution this can be substituted by:

vH v 3 8 v
    122  with: =0.4 and =0.03 (8-68)
2 z so   so

Figure 8-29, Figure 8-30 and Figure 8-31 give the removal ratio or settling efficiency for individual particles for
values of λ of 0.01, 0.02 and 0.03.

T o ta l s e ttl in g e ffic ie n c y fo r i n d iv i d u a l g r a in s ( la b d a = 0 .0 1 ) .
1 .0
G ra in s e t tlin g e ffic ie n c y

v /v o = 0 . 1

0 .9 v /v o = 0 . 2

v /v o = 0 . 3

v /v o = 0 . 4
0 .8 v /v o = 0 . 5

v /v o = 0 . 6

v /v o = 0 . 7
0 .7
v /v o = 0 . 8

v /v o = 0 . 9

0 .6 v /v o = 1 . 0

v /v o = 1 . 1
rb = rg *rt

v /v o = 1 . 2
0 .5 v /v o = 1 . 3

v /v o = 1 . 4

v /v o = 1 . 5
0 .4
v /v o = 1 . 6

v /v o = 1 . 7

v /v o = 1 . 8
0 .3
v /v o = 1 . 9

v /v o = 2 . 0

0 .2 v /v o = 2 . 5

v /v o = 3

v /v o = 3 . 5
0 .1
v /v o = 4

v /v o = 4 . 5

v /v o = 5
0 .0
0 .0 0 0 1 0 .0 0 1 0 .0 1 0 .1 1

v /s o

Figure 8-29: The total settling efficiency for λ=0.01.

Page 246 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

T o ta l s e ttl in g e ffic ie n c y fo r i n d iv i d u a l g r a in s ( la b d a = 0 .0 2 ) .
1 .0
G ra in s e t tlin g e ffic ie n c y

v /v o = 0 . 1

0 .9 v /v o = 0 . 2

v /v o = 0 . 3

v /v o = 0 . 4
0 .8 v /v o = 0 . 5

v /v o = 0 . 6

v /v o = 0 . 7
0 .7
v /v o = 0 . 8

v /v o = 0 . 9

0 .6 v /v o = 1 . 0

v /v o = 1 . 1
rb = rg *rt

v /v o = 1 . 2
0 .5 v /v o = 1 . 3

v /v o = 1 . 4

v /v o = 1 . 5
0 .4
v /v o = 1 . 6

v /v o = 1 . 7

v /v o = 1 . 8
0 .3
v /v o = 1 . 9

v /v o = 2 . 0

0 .2 v /v o = 2 . 5

v /v o = 3

v /v o = 3 . 5
0 .1
v /v o = 4

v /v o = 4 . 5

v /v o = 5
0 .0
0 .0 0 0 1 0 .0 0 1 0 .0 1 0 .1 1

v /s o

Figure 8-30: The total settling efficiency for λ=0.02.

T o ta l s e ttl in g e ffic ie n c y fo r i n d iv i d u a l g r a in s ( la b d a = 0 .0 3 ) .
1 .0
G ra in s e t tlin g e ffic ie n c y

v /v o = 0 . 1

0 .9 v /v o = 0 . 2

v /v o = 0 . 3

v /v o = 0 . 4
0 .8 v /v o = 0 . 5

v /v o = 0 . 6

v /v o = 0 . 7
0 .7
v /v o = 0 . 8

v /v o = 0 . 9

0 .6 v /v o = 1 . 0

v /v o = 1 . 1
rb = rg *rt

v /v o = 1 . 2
0 .5 v /v o = 1 . 3

v /v o = 1 . 4

v /v o = 1 . 5
0 .4
v /v o = 1 . 6

v /v o = 1 . 7

v /v o = 1 . 8
0 .3
v /v o = 1 . 9

v /v o = 2 . 0

0 .2 v /v o = 2 . 5

v /v o = 3

v /v o = 3 . 5
0 .1
v /v o = 4

v /v o = 4 . 5

v /v o = 5
0 .0
0 .0 0 0 1 0 .0 0 1 0 .0 1 0 .1 1

v /s o

Figure 8-31: The total settling efficiency for λ=0.03.

Copyright © Dr.ir. S.A. Miedema TOC Page 247 of 304


Introduction Dredging Engineering.

The settling efficiency for v/vo<1 can be approximated by equation (8-69), while equation (8-70) gives a good
approximation for the case v/vo>1:

    .1 3  .8 0     v   
 .8 8 5  .2 0   g  .3 3  .9 4   g
 t   g   1  .1 8 4   g   1  T anH  g   L og    .2 6 1 4  .5  L o g    g 
0
  
g

    
     so   
(8-69)
    .7 7  .0 8     v     
 .6 9  .3 8   g  1 .0 1  .1 8   g
  .2 6 1 4  .5  L o g   
1
 t   g   1  .1 8 4   g   1  T anH  g   L og  g   
g

  
   
     so     
(8-70)

The effect of turbulence is taken into account by multiplying the settling efficiency with the turbulence efficiency
ηt according to Miedema & Vlasblom (1996). Since the turbulence efficiency is smaller than 1 for all grains
according to the equations (8-69) and (8-70), the basin settling efficiency can be determined with equation (8-71),
where ps equals 0 as long as scour does not occur. So the total settling efficiency is now:

1
b   g  t  dp (8-71)
ps

Page 248 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

8.9. Comparing the Miedema and the van Rhee Models.


8.9.1. Introduction.
This chapter is based on Miedema & van Rhee (2007).
In the past two decades the size of TSHD’s has tripled and there are plans for TSHD’s in the range of 50.000 m 3.
When enlarging hoppers there are some limitations like the draught of the vessel and the line velocity in the suction
lines. It’s interesting to compare the influences of length, width, height ratio’s, flow capacity and some other
parameters on the production and the overflow losses of TSHD’s. To do so, mathematical models have been
developed to simulate the sedimentation process in the hopper. Two models will be used and compared, first the
model of Vlasblom/Miedema (1995), Miedema/Vlasblom (1996) and Miedema (2008A) and second the more
sophisticated 2DV model of van Rhee (2002C), which is verified and validated with model and prototype tests.
Both models are explained briefly. With the two models 3 cases are analyzed, a 2316 m3, a 21579 m3 and a 36842
m3 hopper. The results of the case studies give the following conclusions and recommendations:
 The two models give the same magnitude for the overflow losses, but the shape of the curves is different due
to the differences in the physical modeling of the processes.
 Due to the lower losses the computed optimal loading time will be shorter for the Vlasblom /Miedema
approach.
 The strong point of the van Rhee model is the accurate physical modeling, giving the possibility to model the
geometry of the hopper in great detail, but also describing the physical processes in more detail.
 The van Rhee model is verified and validated with model and prototype tests and can be considered a reference
model for other models.
 The strong point of the Miedema/Vlasblom model is the simplicity, giving a transparent model where result
and cause are easily related.
From a scientific point of view it is interesting to compare the sophisticated van Rhee model with the simplified
models and to do so, the van Rhee (2002C) model is compared with the Miedema (2008A) model. The comparison
consists of a number of cases regarding real TSHD’s. The following TSHD’s will be compared:

Table 8-2: The data of the TSHD's used.


Hopper Load Volume Length Width Empty Flow Hopper Mixture
height load v0 density
ton m3 m m m m3/sec m/sec ton/m3
Small 4400 2316 44.0 11.5 4.577 4 0.0079 1.3
Jumbo 41000 21579 79.2 22.4 12.163 14 0.0079 1.3
Mega 70000 36842 125.0 30.0 9.825 19 0.0051 1.3

Further it is assumed that all 3 TSHD’s have a design density of 1.9 ton/m3 and they operate according to the CVS
system (no adjustable overflow). This gives a sand fraction of 0.54 and a porosity of 0.46. For the calculations a
sand with a d50 of 0.4 mm is chosen, according to figure 1. The particle size distribution is chosen in such a way
that there is a reasonable percentage of fines in order to have moderate overflow losses.

8.9.2. Case Studies with the Camp/Miedema Model.


The calculations according to the modified Camp/Miedema model as developed by Miedema (1981) and published
by Vlasblom & Miedema (1995), Miedema & Vlasblom (1996) and Miedema (2008A) are carried out with the
program TSHD (developed by Miedema). The effects of hindered settling, turbulence and scour and an adjustable
overflow are implemented in this program as described previously.
The program assumes that first the hopper is filled with mixture up to the overflow level and all the grains entering
the hopper during this phase will stay in the hopper, so the overflow losses are 0 during this phase. The table below
shows the filling time, the total load and the TDS at the end of this phase.

Table 8-3: The hopper content after the filling phase.


Hopper Load Volume Flow Filling Total TDS Overflow Mixture
time load losses density
ton m3 m3/sec min ton ton % ton/m3
Small 4400 2316 4 9.65 3011 1039 20.0 1.3
Jumbo 41000 21579 14 25.69 28053 9678 20.0 1.3
Mega 70000 36842 19 32.32 47895 16523 16.6 1.3

Copyright © Dr.ir. S.A. Miedema TOC Page 249 of 304


Introduction Dredging Engineering.

After this phase the program will determine the total settling efficiency and based on this the increase of the
sediment and the overflow losses in time steps of 1 minute. Each time step the program checks whether or not
scour occurs and if so which fraction of the PSD will not settle due to scour. Usually first there is a phase where
scour does not occur. The overflow losses are determined by the settling efficiency according to the equations
(8-69) and (8-70). If the hopper has a CTS system, each time the necessary overflow level is calculated and the
overflow level is adjusted. In the cases considered a CVS system is assumed, so the overflow level is fixed. When
the sediment level is so high that the velocity above the bed is very high, scour starts. This will happen at the end
of the loading process. In the calculations the loading process is continued for a while, so the effect of scour is
clearly visible. The results of the calculations are show in Figure 8-33, Figure 8-34 and Figure 8-35 for the Small,
Jumbo and Mega hopper. The initial overflow losses of 20, 20 and 16.6% match the values of the hopper load
parameter as mentioned in Table 8-2. The Mega hopper has a smaller hopper load parameter and thus also smaller
initial overflow losses (without scour).

100

80
Cum. percentage

60

40

20

0
0.01 0.1 1 10 100
Particle size [m m ]

PSD in Vlasblom/Miedema Northsea Coarse


PSD in 2DV Model

Figure 8-32: The 0.4 mm grain distribution.

T h e lo a d in g c u rv e s .
5 00 0 5 00 0

4 50 0 4 50 0

4 00 0 4 00 0

3 50 0 3 50 0

3 00 0 3 00 0
L o a d in t o n s
L o a d in t o n s

2 50 0 2 50 0

2 00 0 2 00 0

1 50 0 1 50 0

1 00 0 1 00 0

5 00 5 00

0 0
0 .0 6 .0 1 2.0 1 8.0 2 4.0 3 0.0 3 6.0 4 2.0 4 8.0 5 4.0 6 0.0

T im e in m i n

T o ta l lo a d E ff e c t iv e ( s itu ) lo a d T o n n e s D r y S o lid s O v e r flo w lo s s e s e ffe c tiv e O v e r flo w lo s s e s T D S

Figure 8-33: The loading curves of the Small TSHD.

Page 250 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

T h e lo a d in g c u rv e s .
4 50 0 0 4 50 0 0

4 05 0 0 4 05 0 0

3 60 0 0 3 60 0 0

3 15 0 0 3 15 0 0

2 70 0 0 2 70 0 0

L o a d in t o n s
L o a d in t o n s

2 25 0 0 2 25 0 0

1 80 0 0 1 80 0 0

1 35 0 0 1 35 0 0

9 00 0 9 00 0

4 50 0 4 50 0

0 0
0 .0 1 5.0 3 0.0 4 5.0 6 0.0 7 5.0 9 0.0 1 0 5 .0 1 2 0 .0 1 3 5 .0 1 5 0 .0

T im e in m i n

T o ta l lo a d E ff e c t iv e (s itu ) lo a d T o n n e s D ry S o lid s O v e rflo w lo s s e s e ffe c tiv e O v e rflo w lo s s e s T D S

Figure 8-34: The loading curves of the Jumbo TSHD.

T h e lo a d in g c u rv e s .
7 50 0 0 7 50 0 0

6 75 0 0 6 75 0 0

6 00 0 0 6 00 0 0

5 25 0 0 5 25 0 0

4 50 0 0 4 50 0 0

L o a d in t o n s
L o a d in t o n s

3 75 0 0 3 75 0 0

3 00 0 0 3 00 0 0

2 25 0 0 2 25 0 0

1 50 0 0 1 50 0 0

7 50 0 7 50 0

0 0
0 .0 2 5.0 5 0.0 7 5.0 1 0 0 .0 1 2 5 .0 1 5 0 .0 1 7 5 .0

T im e in m i n

T o ta l lo a d E ff e c t iv e (s itu ) lo a d T o n n e s D ry S o lid s O v e rflo w lo s s e s e ffe c tiv e O v e rflo w lo s s e s T D S

Figure 8-35: The loading curves of the Mega TSHD.

It should be noted that the optimum loading time, the loading time with the maximum production, depends on the
total cycle, including sailing times, dumping time, etc. Since the calculations with the 2DV model start with a
hopper full of water, also here first the hopper is filled with water, so the two models can be compared.

Copyright © Dr.ir. S.A. Miedema TOC Page 251 of 304


Introduction Dredging Engineering.

8.9.3. The 2DV Model

The settlement model described above provides a good approximation of the overflow losses. The influence of
grain size, discharge, concentration and hopper geometry can be taken into account. Some influences however are
not included in the model. For instance the influence of the inflow location, variation of water level at the start of
dredging is not included. To overcome these limitations the 2DV hopper sedimentation model was developed (Van
Rhee (2002A)). The model is based on the Reynolds Averaged Navier Stokes equations with a k-epsilon turbulence
model. The model includes the influence of the overflow level of the hopper (moving water surface) and a moving
sand bed due to the filling of the hopper. The influence of the particle size distribution (PSD) is included in the
sediment transport equations. A summary of the model is described in Van Rhee (2002C). The total model is
based on three modules (see Figure 8-36).

Figure 8-36: Overview of the 2DV model.

In the 2D RANS module the Reynolds Averaged Navier Stokes equations are solved (the momentum equations).
The sediment transport module computes the distribution of suspended sediment in the hopper while the k-epsilon
module is necessary for the turbulent closure. The modules have to be solved simultaneously because the equations
are strongly coupled. In the momentum equations the density is present which follows from the sediment transport
equations. The diffusive transport of sediment is governed by turbulence predicted by the k-epsilon model. The
turbulence on the other hand is influenced by the density gradients computed in the sediment transport module.

Boundary conditions
The partial differential equations can be solved in case boundary conditions are prescribed. Different boundaries
can be distinguished: Walls (sediment bed and side walls), water surface, inflow section and outflow section. At
the walls the normal flow velocity is zero. The boundary condition for the flow velocity at the wall is computed
using a so-called wall function (Rodi (1993), Stansby (1997)). The boundary conditions for the turbulent energy
k en dissipation rate ε are consistent with this wall function approach. For the sediment transport equations the
fluxes through vertical walls and water surface is equal to zero since no sediment enters or leaves the domain at
these boundaries. At the sand bed for every fraction the sedimentation flux Si is prescribed (the product of the near
bed concentration and vertical particle velocity of a certain fraction). The influence of the bottom shear stress on
the sedimentation is modeled using a reduction factor R.

Si  R ci  w zj

  (8-72)
1    0
R   0
 0   0

This simple relation between the reduction factor and Shields parameter θ is based on flume tests (Van Rhee
(2002B)). The critical value for the Shields parameter proved to be independent of the grain size for the sands

Page 252 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

tested (d50 < 300 μm). It will be clear that this approach can only be used when overall sedimentation (like in a
hopper of a TSHD) will take place. When the Shields value exceeds the critical value no sedimentation will take
place, but sediment already settled will not be picked up with this approach. Hence net erosion is not (yet) possible
in the model.

At the inflow section the velocity and concentration is prescribed. The outflow boundary is only active when
overflow is present, so when the mixture level in the hopper exceeds the overflow level. In that case the outflow
velocity is prescribed, and follows simply from the ratio of the overflow discharge and the difference between the
hopper and overflow level. For the other quantities the normal gradients are equal to zero (Neumann condition).
At the water surface a rigid-lid assumption is used since surface wave phenomena are not important for the subject
situation. A rigid-lid can be regarded as a smooth horizontal plate covering the water surface in the hopper.
Depending on the total volume balance inside the hopper this “plate” will be moved up and down.

Numerical approach
The momentum and sediment transport equations are solved using the Finite Volume Method to ensure
conservation. The transport equations for the turbulent quantities k and are solved using the Finite Difference
method. A Finite Difference Method is allays implemented on a rectangular (Cartesian) grid. Although a Finite
Volume Method can be applied on any grid it is advantageous to use a Cartesian approach for this method as well
especially when a staggered arrangement of variables is used. In general the flow domain is however not
rectangular. The water surface can be considered horizontal on the length scale considered, but a sloping bottom
will not coincide with the gridlines. Different approaches are possible. The first method is to use a Cartesian grid
and to adjust the bottom cells (cut-cell method). Another method is to fit the grid at the bottom. In that case a
boundary fitted non-orthogonal grid can be used. A third method is using grid transformation. By choosing an
appropriate transformation the equations are solved on a Cartesian domain in transformed co-ordinates. Although
this transformation allows for a good representation of a curved topography the method has the disadvantage that
due to truncation errors in the horizontal momentum equation artificial flows will develop when a steep bottom
encounters density gradients. These unrealistic flows can be partly suppressed when the diffusion terms are locally
discretized in a Cartesian grid (Stelling (1994)). Since however in a hopper both large density gradients as steep
bottom geometry can be present it was decided to develop the model in Cartesian co-ordinates with a cut-cell
approach at the bed.
The computational procedure can only be outlined here very roughly. The flow is not stationary hence the system
is evaluated in time. The following steps are repeated during time:
 Update the velocity field to time tn+1 by solving the NS-equations together with the continuity equation
using a pressure correction method (SIMPLE-method (Patankar (1980)) using the density and eddy
viscosity of the old time step tn.
 Update the turbulent quantities and to time tn+1 using the velocity field of tn+1. Compute the eddy-
viscosity for the new time.
 Use the flow field of tn+1 to compute the grain velocities for the next time and update the concentrations
for all fractions and hence the mixture density to time tn+1.
 Compute the new location for the bed level and mixture surface in the hopper

Results
The 2DV model is used to simulate the loading process for the three different cases. At the start of the simulation
the hopper is filled with water. The results are shown in Figure 8-37, Figure 8-38 and Figure 8-39. In these figures
the TDS in the hopper (settled in the bed and in suspension) and the cumulative overflow losses are plotted versus
loading time.

Copyright © Dr.ir. S.A. Miedema TOC Page 253 of 304


Introduction Dredging Engineering.

Small size TSHD


3,500

3,000

2,500

2,000

1,500

1,000

500

0
0 6 12 18 24 30 36 42
time [min]

TDS in hopper [ton] overflow loss [ton]

Figure 8-37: Loaded TDS and overflow losses as a function of time for a Small size TSHD.

Jumbo TSHD
36000

31500

27000

22500

18000

13500

9000

4500

0
0 15 30 45 60 75 90 105 120
time [min]

TDS in hopper [ton] overflow loss [ton]

Figure 8-38: Loaded TDS and overflow losses as a function of time for Jumbo TSHD.

Page 254 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

Mega TSHD
60000

52500

45000

37500

30000

22500

15000

7500

0
0 25 50 75 100 125 150
time [min]

TDS in hopper [ton] overflow loss [ton]

Figure 8-39: Loaded TDS and overflow losses as a function of time for a mega TSHD.

8.9.4. Comparison of the Two Models.

To compare the results of the two methods, first the differences in the models are summarized:
1. The physical modeling of the two methods is different; Miedema/Vlasblom/Camp is based on the Camp
approach, while the 2DV model is based on the Reynolds Averaged Navier Stokes equations.
2. The van Rhee model starts with a hopper full of water, while the Miedema/Vlasblom/Camp model starts
with an empty hopper.
3. The Miedema/Vlasblom/Camp model assumes 100% settling of the grains during the filling phase of the
hopper.
4. The van Rhee model includes a layer of water above the overflow level, while the
Miedema/Vlasblom/Camp model doesn’t by default. But to compare the two models the height of the
overflow level has been increased by the thickness of this layer of water and the results are shown in the
2/3
 Q 
Figure 8-40, Figure 8-41 and Figure 8-42. With the layer thickness according to: Hl    ,
 1 . 72  b 
where the constant 1.72 may vary. The width W is chosen for the width of the overflow b in the
calculations. This gives a layer thickness of 34 cm for the small hopper and 51 cm for the Jumbo and the
Mega hopper.

The results of the Small hopper and the Jumbo hopper are similar due to the same hopper load parameter of 0.0079
m/sec. The Mega hopper has a smaller hopper load parameter of 0.0051 m/sec, resulting in relatively smaller
overflow losses. To compare the two models the graphs of the two models are combined and similarities and
differences are discussed:

Similarities:
1. The overflow rate seems to be quite similar for all 3 hoppers, until the Miedema/Vlasblom/Camp
approach reaches the scour phase. From this moment on the overflow rate increases rapidly.
2. It is obvious that at the end of the loading both models find the same amount of sand in all cases, since
this matches the maximum loading capacity of the hopper in question. This observation explains the fact
that the overflow losses of both models are almost the same at the time where the van Rhee simulation

Copyright © Dr.ir. S.A. Miedema TOC Page 255 of 304


Introduction Dredging Engineering.

stops (42 minutes for the Small hopper, 112 minutes for the Jumbo hopper and 137 minutes for the Mega
hopper).

Differences:
1. The overflow losses in the van Rhee model are lower in the first phase, because in the
Miedema/Vlasblom/Camp approach this occurs instantly, while the van Rhee approach considers the time
the mixture needs to flow through the hopper and the effect of scour is very limited because a uniform
flow velocity distribution over depth is assumed (leading to very low horizontal flow velocities) in this
model. Only at the end of the loading stage the effect of the horizontal flow velocity on sedimentation
becomes noticeable. For instance for the Small hopper the TDS loading curve is a straight line from the
start of overflow up to 33 min after start dredging. After that time the loading rate decreases as a result of
the increasing horizontal velocity. At t = 45 min the hopper is completely filled. Hence the influence of
the velocity during the final loading stage is present for about 12 minutes.
2. In the 2DV model velocity distribution is not prescribed, but is determined by physics and depends on
the inflow conditions. In general, due to the large density difference between the inflowing mixture and
fluid already present in the hopper, density currents will develop. This will lead to a larger velocity close
to the sand bed surface. Hence the effect of the flow velocity on sedimentation will be present from the
start of dredging. This influence does not increase much during loading. The effect is more spread out
over the loading cycle. The loading rate decreases gradually, but remains on a reasonable level unto the
moment that the hopper is fully loaded. In the Miedema/Vlasblom/Camp loading rate reduces to zero at
full load..
3. If optimum loading time is considered, the two models differ in that the van Rhee model gives 43, 112
and 137 minutes, while this will be around 38, 99 and 120 minutes in the Miedema/Vlasblom/Camp
approach. Both models start with a hopper full of water, so this should be considered. The overflow losses
in the final phase of the loading process are similar for both models.

T h e lo a d in g c u rv e s .
4 00 0 4 00 0

3 60 0 3 60 0

3 20 0 3 20 0

2 80 0 2 80 0

2 40 0 2 40 0

L o a d in t o n s
L o a d in t o n s

2 00 0 2 00 0

1 60 0 1 60 0

1 20 0 1 20 0

8 00 8 00

4 00 4 00

0 0
0 .0 2 .5 5 .0 7 .5 1 0.0 1 2.5 1 5.0 1 7.5 2 0.0 2 2.5 2 5.0 2 7.5 3 0.0 3 2.5 3 5.0 3 7.5 4 0.0 4 2.5

T im e in m i n
L o a d T D S M ie d e m a O v e rflo w T D S M ie d e m a L oa d TD S va n R he e O v e rflo w T D S v a n R h e e T D S in M ie d e m a T D S in v a n R h e e

Figure 8-40: Comparison of the two models for the Small hopper.

Page 256 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

T h e lo a d in g c u rv e s .
3 60 0 0 3 60 0 0

3 30 0 0 3 30 0 0

3 00 0 0 3 00 0 0

2 70 0 0 2 70 0 0

2 40 0 0 2 40 0 0

2 10 0 0 2 10 0 0

L o a d in t o n s
L o a d in t o n s

1 80 0 0 1 80 0 0

1 50 0 0 1 50 0 0

1 20 0 0 1 20 0 0

9 00 0 9 00 0

6 00 0 6 00 0

3 00 0 3 00 0

0 0
0 .0 1 2.5 2 5.0 3 7.5 5 0.0 6 2.5 7 5.0 8 7.5 1 0 0 .0 1 1 2 .5

T im e in m i n
L o a d T D S M ie d e m a O v e rflo w T D S M ie d e m a L oa d TD S va n R he e O v e rflo w T D S v a n R h e e T D S in M ie d e m a T D S in v a n R h e e

Figure 8-41: Comparison of the two models for the Jumbo hopper.

T h e lo a d in g c u rv e s .
6 00 0 0 6 00 0 0

5 50 0 0 5 50 0 0

5 00 0 0 5 00 0 0

4 50 0 0 4 50 0 0

4 00 0 0 4 00 0 0

3 50 0 0 3 50 0 0
L o a d in t o n s
L o a d in t o n s

3 00 0 0 3 00 0 0

2 50 0 0 2 50 0 0

2 00 0 0 2 00 0 0

1 50 0 0 1 50 0 0

1 00 0 0 1 00 0 0

5 00 0 5 00 0

0 0
0 .0 1 2.5 2 5.0 3 7.5 5 0.0 6 2.5 7 5.0 8 7.5 1 0 0 .0 1 1 2 .5 1 2 5 .0 1 3 7 .5

T im e in m i n
L o a d T D S M ie d e m a O v e rflo w T D S M ie d e m a L oa d TD S va n R he e O v e rflo w T D S v a n R h e e T D S in M ie d e m a T D S in v a n R h e e

Figure 8-42: Comparison of the two models for the Mega hopper.

Copyright © Dr.ir. S.A. Miedema TOC Page 257 of 304


Introduction Dredging Engineering.

8.9.5. Conclusions

 The two models give the same magnitude for the overflow losses, but the shapes of the curves are different
due to the differences in the physical modeling of the processes.
 Due to the lower losses the computed optimal loading time will be shorter for the Miedema/Vlasblom /Camp
approach.
 The strong point of the van Rhee model is the accurate physical modeling, giving the possibility to model the
geometry of the hopper in great detail, but also describing the physical processes in more detail.
 The van Rhee model is verified and validated with model and prototype tests and can be considered a reference
model for other models.
 The strong point of the Miedema/Vlasblom/Camp model is the simplicity, giving a transparent model where
result and cause are easily related.

Page 258 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

8.10. A Sensitivity Analysis of the Scaling of TSHS’s.


The loading process of TSHD’s contains a number of non-linearity’s:
1. The real hopper load parameter will vary during the loading process.
2. The turbulence settling efficiency.
3. The behavior of the layer of water above the overflow.
4. The behavior of hindered settling.
5. The effective concentration in the hopper.
6. The so called storage effect.
Based on all these non-linearity’s it is not expected that TSHD’s can be scaled easily, however the research in this
paper shows that with the right choice of scale laws the TSHD’s can be scaled rather well.
4 TSHD’s are chosen, derived from Miedema & van Rhee (2007), but adapted to the scale laws. With each of
these TSHD’s simulations are carried out in 4 types of sand, 400 µm, 250 µm, 150 µm and 100 µm sand.

8.10.1. Scale Laws.


To compare TSHD’s of different dimensions scale laws have to be applied in order to create identical loading
processes. Scale laws should be based on the physical and the operational processes that occur. Further the shape
of the hopper should be identical and the relation with the flow should match. It is however also important to
decide which parameter or parameters to choose for the comparison of the TSHD’s. When can the conclusion be
drawn that two hoppers with different dimensions behave identical. The main parameter that is chosen for this
comparison are the cumulative overflow losses. The cumulative overflow losses are the overflow losses expressed
as TDS (Tonnes Dry Solids) divided by the total amount of TDS that has entered the hopper, from the start of the
loading process until the moment of optimum loading.
The first important parameter to consider is the hopper load parameter (HLP) as described in equation (8-73). Here
the hopper load parameter without the effect of the bed rise velocity is considered, because the bed rise velocity
changes during the loading process and would result in changing scale laws. As stated before, the hopper load
parameter is the settling velocity of a grain that will settle for 100%. Larger grains will also settle for 100%, but
smaller grains will settle with a smaller percentage.

H w
Q in
vo  so   (8-73)
L W L

If two TSHD’s with different dimensions have the same hopper load parameter, it can be expected that under
similar conditions, the momentary overflow losses are equal and thus also the cumulative overflow losses.
However the hopper load parameter does not take into consideration the effects of turbulence efficiency, hindered
settling, and the storage effect and so on.

A second scale law could be that the ratios between Length, Width and Height are identical. If a length scale λ is
considered this gives:

L1 W1 H1 H L P1 Q1 2 Tf 1 V1 / Q 1
    and  1 and   and    (8-74)
L2 W2 H 2
H L P2 Q 2
Tf 2 V2 / Q 2

Because the hopper load parameter is considered to be a constant, the flow Q will scale with the square of the
length scale λ. The filling time Tf, which is the time to fill the hopper up to the overflow level also scales with the
length scale λ. To have similar processes for determining the optimum loading time, the travelling time, which is
the sum of the sailing time to and from the dump area and the dumping time, should also be scaled with the length
scale, assuming that the loading time is proportional to the filling time. Since the horizontal flow velocity in the
hopper equals the flow Q divided by the width W and the height H of the hopper, the horizontal flow velocity is
a constant and does not depend on the length scale. This also follows from the fact that the hopper load parameter
is a constant. If it is assumed that the maximum line velocity in the suction pipes is a constant, for example 7 m/s
and because the line velocity equals the flow velocity divided by 2 and divided by the cross section of one pipe,
this implies that the pipe diameter should be proportional to the square root of the flow and thus be proportional
to the length scale λ.
Because sand is difficult to scale and in reality the sand will be the same independent of the TSHD used, it is
assumed that the sand is the same for all hopper sizes. This implies that the settling velocities are the same and
looking at the equations (8-69) and (8-70) this means that the grain settling efficiency ηg does not depend on the

Copyright © Dr.ir. S.A. Miedema TOC Page 259 of 304


Introduction Dredging Engineering.

hopper size and the ratio vs/so does not depend on the hopper size, since the horizontal flow velocity so does not
depend on the hopper size. The resulting turbulence efficiency as calculated with equations (8-69) and (8-70) is
thus not dependent on the hopper size, although it will change during the loading process.

8.10.2. The TSHD’S used.


Based on the scale laws and based on Miedema & van Rhee (2007), 4 TSHD’s are chosen in a range from small
to Mega. The main dimensions and additional parameters of these hoppers can be found in table 1 and 2.

Table 8-4: The main dimensions of the 4 TSHD's.


Hopper Length (m) Width (m) Empty Volume Design Maximum HLP
height (m) (m3) density load (ton) (m/sec)
(ton/m3)
Small 40 10 5.0 2000 1.5 3000 0.008
Large 60 15 7.5 6750 1.5 10125 0.008
Jumbo 80 20 10.0 16000 1.5 24000 0.008
Mega 100 25 12.5 31250 1.5 46875 0.008

Table 8-5: Additional and derived quantities.


Hopper Flow Pipe Filling Sailing Hydraulic Reynolds Mixture
(m3/sec) diameter time (min) time (min) diameter number density
(m) (m) (ton/m3)
Small 3.2 0.54 10.4 104 10 0.64*106 1.3
Large 7.2 0.81 15.6 156 15 0.96*106 1.3
Jumbo 12.8 1.08 20.8 208 20 1.28*106 1.3
Mega 20.0 1.35 26.0 260 25 1.60*106 1.3

Table 8-4 and Table 8-5 show a wide range of TSHD’s from Small (2000 m3) to Mega (31250 m3) . As can be
noted in the tables, the hopper load parameters are constant at 0.008 m/sec, which is the settling velocity of a grain
a bit bigger than 100 µm. The design density of the TSHD’s is chosen at 1.5 ton/m3, which implies that the loading
process will follow the Constant Tonnage Loading process. The total sailing and dumping time is chosen 10 times
the filling time, which of course is arbitrary, but the resulting sailing times seem to be representative for the reality.
The mixture density is chosen at 1.3 ton/m3, which is high enough to take the influence of hindered settling into
account. It should be noted that the Reynolds numbers of the horizontal flow in the hopper are not constant; the
Reynolds numbers are proportional to the length scale λ. The question is whether or not this will influence the
loading process. As stated before, it does not influence the turbulent settling efficiency, but it could influence the
scour in the final phase of the loading process. Scour is influenced by the viscous friction of the fluid flowing over
the bed. This friction depends on the relative roughness and the Reynolds number. The roughness of the sediment
has the magnitude of the grain diameter which is in the range of 0.1-0.5 mm, while the hydraulic diameters of the
4 TSHD’s are in the magnitude of 10-25 m. The largest relative roughness would occur for a 0.5 mm grain and a
hydraulic diameter of 10 m, giving 0.0005/10=0.00005. The friction coefficient will be between 0.0175 and
0.0171, which hardly has an effect on the scour. Although there will always be some effect, it is not expected that
this effect will have a big influence on the similarity of the loading processes of the 4 TSHD’s. The sediment
density is chosen at 1.9 ton/m3, which means that the TDS is about 76% of the weight of the wet sediment.

For carrying out the simulations 4 grain distributions are chosen. All 4 grain distributions have a d15 for grains
with a settling velocity smaller than the hopper load parameter in order to be sure there will be significant overflow
losses. If grain distributions were chosen with almost 100% of the grains having a settling velocity above the
hopper load parameter, this would result in very small cumulative overflow losses and a good comparison would
be difficult. Table 8-6 gives the d15, d50 and d85 of the 4 grain distributions, while figure 12 shows the full PSD’s.

Table 8-6: The characteristics of the 4 grain distributions.


400 µm 250 µm 150 µm 100 µm
d15 70 µm 80 µm 80 µm 50 µm
d50 400 µm 250 µm 150 µm 100 µm
d85 2000 µm 750 µm 300 µm 200 µm

Page 260 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

C u m u la tiv e G ra in S iz e D is trib u tio n


100

90

80
F in e r b y W e ig h t

70

60

50

40

30
%

20

10

0
0 .0 0 1 0 .0 1 0 .1 1 10 100 1000
G ra in S iz e in m m
0 .1 0 0 m m 0 .1 5 0 m m 0 .2 5 0 m m 0 .4 0 0 m m

V. F in e F in e M ed iu m Co ars e V. F in e F in e M ed iu m Co ars e V. C o ar s e G r a in s Pe b b les C o b b le s B o u ld e r s

C la y
S ilt Sand G ravel

Figure 8-43: The 4 grain distributions.

8.10.3. Simulation Results.


The simulations of the loading process of the 4 TSHD’s are carried out with software based on the model published
by Miedema (2008A), including turbulence efficiency, hindered settling, the storage effect, the layer of water
above the overflow and more. The results of these simulations are summarized in Table 8-7, Table 8-8, Table 8-9
and Table 8-10.

Table 8-7: The simulation results with the 0.400 mm sand.


400 µm sand Loading time TDS (ton) Overflow losses Cumulative Production
(min) TDS (ton) overflow losses (%) (ton/min)
Small 31.0 2174 476 18.0% 16.1
Large 46.5 7349 1594 17.8% 36.2
Jumbo 62.0 17440 3758 17.7% 64.5
Mega 77.5 34089 7313 17.7% 100.9

Table 8-8: The simulation results with the 0.250 mm sand.


250 µm sand Loading time TDS (ton) Overflow losses Cumulative Production
(min) TDS (ton) overflow losses (%) (ton/min)
Small 31.0 2146 503 19.0% 15.9
Large 46.5 7258 1685 18.8% 35.8
Jumbo 61.8 17218 3923 18.6% 63.7
Mega 77.3 33662 7651 18.5% 99.7

Table 8-9: The simulation results with the 0.150 mm sand.


150 µm sand Loading time TDS (ton) Overflow losses Cumulative Production
(min) TDS (ton) overflow losses (%) (ton/min)
Small 32.2 2104 645 23.5% 15.4
Large 48.2 7114 2149 23.2% 34.8
Jumbo 64.2 16887 3923 23.0% 62.0
Mega 80.3 33030 7651 23.0% 96.9

Copyright © Dr.ir. S.A. Miedema TOC Page 261 of 304


Introduction Dredging Engineering.

Table 8-10: The simulation results with the 0.100 mm sand.


100 µm sand Loading time TDS (ton) Overflow losses Cumulative Production
(min) TDS (ton) overflow losses (ton/min)
(%)
Small 43.0 2111 1564 42.6% 14.3
Large 64.7 7145 5292 42.6% 32.3
Jumbo 86.0 16952 12452 42.3% 57.6
Mega 107.7 33149 24368 42.4% 90.1

To visualize the simulations, the graphs of the simulations of the Small TSHD and the Mega TSHD can be found
in the Figure 8-44, Figure 8-45, Figure 8-46, Figure 8-47, Figure 8-48, Figure 8-49, Figure 8-50 and Figure 8-51.
From these graphs and the above tables it will be clear that the cumulative overflow losses do not depend on the
size of the TSHD in quantity and in shape op de loading and overflow curves. To understand the above tables and
the following figures, they will be explained and discussed each.

Table 8-7, Table 8-8, Table 8-9 and Table 8-10 show the loading times in the second column, it is clear that the
loading times are almost proportional to the length scale λ and they increase with increasing overflow losses. The
finer the sand, the longer the loading time. The third column gives the TDS at the point of optimum loading. The
TDS of a hopper filled with sediment is about 76% of the weight of the sediment, but since there is still some water
on top of the sediment at the moment of optimum loading the TDS is a bit less. This means that the maximum
TDS of the Small TSHD is 2280 tons, for the Large TSHD 7695 tons, for the Jumbo TSHD 18240 tons and for
the Mega TSHD 35625 tons, so the assumption is correct. The TDS does not depend on the type of sand. The
fourth column gives the overflow losses in tons TDS. Again TDS means, only the weight of the solids, excluding
the pore water and the water on top of the sediment. The fifth column gives the cumulative overflow losses, which
are almost constant for each type of sand. For the 400 µm sand about 17.8%, for the 250 µm about 18.7%, for the
150 µm sand about 23.2% and for the 100 µm sand about 42.4%. These cumulative overflow losses are the
overflow losses in TDS, divided by the total amount of TDS that has entered the hopper. It is clear that the
cumulative overflow losses do not seem to depend on the size of the TSHD, given the scale laws applied in the
simulations. Apparently the scale laws applied are the correct scale laws for scaling TSHD’s in order to get similar
loading and sedimentation processes. It is interesting however to compare the cumulative overflow losses with the
grain size distribution curves of the sands used. The hopper load parameter of 0.008 m/s matches a grain with a
diameter of 0.112 mm. If the percentage of grains smaller than this diameter is considered and compared we the
overflow losses, the following numbers are found. For the 400 µm sand, about 20% smaller than 0.112 mm and
cumulative overflow losses of 17.8%, for the 250 µm sand, about 20% smaller than 0.112 mm and 18.7%
cumulative overflow losses, for the 150 µm sand, about 26% smaller than 0.112 mm and 23.2% cumulative
overflow losses and for the 100 µm sand, about 52% smaller than 0.112 mm and 42.4% cumulative overflow
losses. Apparently, but not unexpected, the cumulative overflow losses have a strong relation with the percentage
of the grains smaller than the grain diameter matching the hopper load parameter. There is however not a fixed
relation, because the grains smaller than the diameter matching the hopper load parameter will still settle partially
and this depends strongly on the steepness of the cumulative grain size distribution. In the examples given it is
clear that the 400 µm sand and the 250 µm sand, both have about 20% smaller and both have a cumulative overflow
loss of about 20%. The simulations however also take hindered settling, the effect of the concentration on the
settling velocity, into account and in reality the TSHD might make turns, resulting in a more complicated loading
process. The overflow losses will also depend on the concentration as will be discussed later. The last column
shows the production and of course the production is decreasing if the cumulative overflow losses are increasing.

Figure 8-44 and Figure 8-45 give the loading curves of the Small and the Mega TSHD in order to see if not only
the cumulative overflow losses are independent of the size of the TSHD, but also the shape of the loading curves.
To understand these graphs the different curves are explained. The loading process starts with an empty hopper,
so there is no water in the hopper. First for 10.4 minutes for the Small hopper and 26.0 minutes for the Mega
hopper, the hopper is filled with mixture of 1.3 ton/m3. After that the loading continues until after about 22.4
minutes for the Small hopper and 57 minutes for the Mega hopper, the maximum load is reached as can be found
in table 1, seventh column. After reaching the maximum load, the loading continues while the overflow is lowered
in such a way that the total load in the hopper remains constant, replacing water above the sediment with sediment.
After about 40 minutes for the Small hopper and about 100 minutes for the Mega hopper, the sediment level is so
high and the layer of water above the sediment is so thin, that very high flow velocities occur above the sediment,
preventing the grains the settle and resulting in scour. After a short while hardly any grains will settle and the
optimum loading point is reached. Continuing after this point will result in a decrease of production and is thus
useless.

Page 262 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

The black solid line at the top is the total load in the hopper and it is obvious that this line stays at the maximum
load once this is reached. The blue solid line is the total volume in the hopper, it can be seen that after reaching
the maximum load, the total volume is decreasing because the overflow is lowered. The dashed red line shows the
tangent method to determine the optimum loading point. The dashed brown line shows the weight of the sediment
in the hopper, including the weight of the pore water. At the end of the loading this line is just below the maximum
load line, because there is still a layer of water above the sediment, which does not count in the sediment weight.
The black solid straight line gives the amount of TDS that enters the hopper, so the sum of sediment TDS and
overflow TDS should be equal to this line. The highest solid brown line is the amount of TDS in the hopper, while
the lowest solid brown line is the sediment volume. Finally the solid red line gives the overflow losses in TDS. It
can be seen that until the mixture in the hopper reaches the overflow level, there are no overflow losses. After the
hopper is filled the overflow losses follow an almost straight line, which curves to a steeper line when scour starts
to occur.
Although the scales of Figure 8-44 and Figure 8-45 are different, it is clear that the different loading curves have
similar shapes, so not only the cumulative overflow losses are independent of the size of the hopper, also the
momentary overflow losses are.

Figure 8-46 and Figure 8-47 show the loading curves including the storage effect. So what exactly is this storage
effect? When grains enter the hopper, it can already be calculated which fraction of the grains will settle and which
fraction of the grains will leave the hopper through the overflow. Figure 8-44 and Figure 8-45 are based on such a
calculation. Grains that will leave through the overflow however, first have to travel through the hopper before
they actually leave the hopper through the overflow. One can say that these grains are temporary stored in the
hopper, the so called storage effect. This means that if suddenly the loading process would stop before the optimum
is reached, there are more grains and thus TDS in the hopper then would follow from the Figure 8-44 and Figure
8-45. It also means that the overflow losses at such a moment would be less. The amount of grains that will leave
the hopper, but are still inside, depends on the time it takes for a particle to move from the entrance to the overflow
and this depends on the flow velocity. The flow velocity will increase when the sediment level increases and at the
end of the loading cycle this velocity is so high that the storage effect can be neglected. In the Figure 8-46 and
Figure 8-47 the top thick solid black lines show the amount of TDS in the hopper (compare with Figure 8-44 and
Figure 8-45, these contain the same lines but solid brown). Just above the thick solid black lines are the thin solid
green lines. The difference between the thick solid black line and the thin solid green line is the amount of TDS
that will leave through the overflow, but has not yet left. The thin solid brown line below the thick solid black line
show how many grains have already settled, the difference between the two lines is the amount of grains that will
settle, but has not yet settled. Finally the thick solid black line at the bottom gives the overflow losses as have
already been shown in Figure 8-44 and Figure 8-45. The thin red line, below this line give the amount of TDS that
have already left the hopper.

Figure 8-48 and Figure 8-49 show the grain distribution curves of the 100 µm for the Small and the Mega TSHD.
The original distribution is the lines with the dots. Left from these are the red lines which give the distribution of
the grains leaving the overflow, on average from the start of the loading until the optimum loading point. Right
from the original distribution is the solid green line, showing the average distribution in the hopper. It can be
concluded that the grain distributions are similar for the Small and the Mega TSHD.

Figure 8-50 and Figure 8-51 show the influence of the concentration and the amount of water in the hopper at the
moment the loading starts, on the cumulative overflow losses and the cumulative efficiency. The dot in both graphs
shows the result of the simulation carried out. It is obvious that Figure 8-50 and Figure 8-51 show similar graphs.
The lines in the graphs are determined by an equation, derived as an attempt to predict the overflow losses with
just one equation. The green solid line shows the cumulative overflow losses when the hopper is completely empty
at the start of the loading process. The blue line when the hopper is filled with 50% water and the red line when
its filled with 100% water. The graph shows the overflow losses as a function of the mixture concentration. These
graphs are still experimental, but give good tendencies of the overflow losses.

Copyright © Dr.ir. S.A. Miedema TOC Page 263 of 304


Introduction Dredging Engineering.

8.10.4. Conclusions & Discussion.


The question before this research started, was how do the cumulative overflow losses behave when TSHD’s are
scaled from small to very big. The second question was, are that scale laws that should be applied when scaling
TSHD’s in order to create similar or maybe even identical processes.

First the answer on the second question, there are scale laws that should be applied and the main law is, to keep
the hopper load parameter constant and from there derive the scale laws for the flow and other dimensions, but
don’t scale the sand.
If the scale laws are applied correctly, the simulations show that scaling the TSHD has hardly any influence on the
cumulative overflow losses and the loading processes are similar.
The overflow losses however depend strongly on the position of the grain diameter match the hopper load
parameter in the particle size distribution diagram. The fraction of the sand with diameters smaller than this
diameter has a very strong relation with the cumulative overflow losses.

Page 264 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

T h e lo a d in g c u rv e s e x c lu d in g th e s to ra g e e ffe c t
3000 3000

2700 2700

2400 2400

2100 2100

W e ig h t ( to n s )
W e ig h t ( to n s )

1800 1800

1500 1500

1200 1200

900 900

600 600

300 300

0 0
0 .0 2 .5 5 .0 7 .5 1 0 .0 1 2 .5 1 5 .0 1 7 .5 2 0 .0 2 2 .5 2 5 .0 2 7 .5 3 0 .0 3 2 .5 3 5 .0 3 7 .5 4 0 .0 4 2 .5 4 5 .0 4 7 .5 5 0 .0

T im e (m in )
Total Weight Total Volume Sediment Weight Tonnes Dry Solids Overflow TDS Dredged TDS Optimum Production Sediment Volume

T ra ilin g S u c tio n H o p p e r D re d g e V 1 .3 , A p ril 2 9 , 2 0 0 9 , 1 6 :1 9 :2 3


S m a ll T S H D ( C :\P r o g r a m F ile s \T r a ilin g S u c tio n H o p p e r D re d g e \T S H D \0 9 S m a ll.In p )
V e ry fin e s a n d d 5 0 = 0 .1 m m ( C :\P r o g r a m F ile s \T r a ilin g S u c tio n H o p p e r D re d g e \S a n d \S a n d .In p )
O p tim u m p ro d u c tio n : 2 1 1 1 T D S , lo a d e d in : 4 3 .0 m in , o v e rflo w lo s s e s : 1 5 6 4 T D S

Figure 8-44: The loading curves for the Small TSHD.

T h e lo a d in g c u rv e s e x c lu d in g th e s to ra g e e ffe c t
50000 50000

45000 45000

40000 40000

35000 35000

W e ig h t ( to n s )
W e ig h t ( to n s )

30000 30000

25000 25000

20000 20000

15000 15000

10000 10000

5000 5000

0 0
0 .0 1 0 .0 2 0 .0 3 0 .0 4 0 .0 5 0 .0 6 0 .0 7 0 .0 8 0 .0 9 0 .0 1 0 0 .0 1 1 0 .0 1 2 0 .0 1 3 0 .0 1 4 0 .0 1 5 0 .0 1 6 0 .0 1 7 0 .0 1 8 0 .0 1 9 0 .0 2 0 0 .0

T im e (m in )
Total Weight Total Volume Sediment Weight Tonnes Dry Solids Overflow TDS Dredged TDS Optimum Production Sediment Volume

T ra ilin g S u c tio n H o p p e r D re d g e V 1 .3 , A p ril 2 9 , 2 0 0 9 , 1 5 :1 1 :2 5


M e g a T S H D (C :\P ro g ra m F ile s \T ra ilin g S u c tio n H o p p e r D r e d g e \T S H D \0 9 M e g a .In p )
V e ry fin e s a n d d 5 0 = 0 .1 m m ( C :\P r o g r a m F ile s \T r a ilin g S u c tio n H o p p e r D re d g e \S a n d \S a n d .In p )
O p tim u m p ro d u c tio n : 3 3 1 4 9 T D S , lo a d e d in : 1 0 7 .7 m in , o v e r flo w lo s s e s : 2 4 3 6 8 T D S

Figure 8-45: The loading curves for the Mega TSHD.

Copyright © Dr.ir. S.A. Miedema TOC Page 265 of 304


Introduction Dredging Engineering.

T h e lo a d in g c u rv e s in c lu d in g th e s to ra g e e ffe c t
3000 3000

2700 2700

2400 2400

2100 2100

W e ig h t ( to n s )
W e ig h t ( to n s )

1800 1800

1500 1500

1200 1200

900 900

600 600

300 300

0 0
0 .0 2 .5 5 .0 7 .5 1 0 .0 1 2 .5 1 5 .0 1 7 .5 2 0 .0 2 2 .5 2 5 .0 2 7 .5 3 0 .0 3 2 .5 3 5 .0 3 7 .5 4 0 .0 4 2 .5 4 5 .0 4 7 .5 5 0 .0

T im e (m in )
O p tim u m T D S P ro d u c t io n O v e rflo w T D S T o ta l T D S in h o p p e r S e d im e n t T D S D re d g e d T D S

T ra ilin g S u c tio n H o p p e r D re d g e V 1 .3 , A p ril 2 9 , 2 0 0 9 , 1 6 :1 9 :2 3


S m a ll T S H D ( C :\P r o g r a m F ile s \T r a ilin g S u c tio n H o p p e r D re d g e \T S H D \0 9 S m a ll.In p )
V e ry fin e s a n d d 5 0 = 0 .1 m m ( C :\P r o g r a m F ile s \T r a ilin g S u c tio n H o p p e r D re d g e \S a n d \S a n d .In p )
O p tim u m p ro d u c tio n : 2 1 1 1 T D S , lo a d e d in : 4 3 .0 m in , o v e rflo w lo s s e s : 1 5 6 4 T D S

Figure 8-46: The loading curves including the storage effect for the Small TSHD.

T h e lo a d in g c u rv e s in c lu d in g th e s to ra g e e ffe c t
50000 50000

45000 45000

40000 40000

35000 35000

W e ig h t ( to n s )
W e ig h t ( to n s )

30000 30000

25000 25000

20000 20000

15000 15000

10000 10000

5000 5000

0 0
0 .0 1 0 .0 2 0 .0 3 0 .0 4 0 .0 5 0 .0 6 0 .0 7 0 .0 8 0 .0 9 0 .0 1 0 0 .0 1 1 0 .0 1 2 0 .0 1 3 0 .0 1 4 0 .0 1 5 0 .0 1 6 0 .0 1 7 0 .0 1 8 0 .0 1 9 0 .0 2 0 0 .0

T im e (m in )
O p tim u m T D S P ro d u c t io n O v e rflo w T D S T o ta l T D S in h o p p e r S e d im e n t T D S D re d g e d T D S

T ra ilin g S u c tio n H o p p e r D re d g e V 1 .3 , A p ril 2 9 , 2 0 0 9 , 1 5 :1 1 :2 5


M e g a T S H D (C :\P ro g ra m F ile s \T ra ilin g S u c tio n H o p p e r D r e d g e \T S H D \0 9 M e g a .In p )
V e ry fin e s a n d d 5 0 = 0 .1 m m ( C :\P r o g r a m F ile s \T r a ilin g S u c tio n H o p p e r D re d g e \S a n d \S a n d .In p )
O p tim u m p ro d u c tio n : 3 3 1 4 9 T D S , lo a d e d in : 1 0 7 .7 m in , o v e r flo w lo s s e s : 2 4 3 6 8 T D S

Figure 8-47: The loading curves including the storage effect for the Mega TSHD.

Page 266 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

C u m u la tiv e G ra in S iz e D is trib u tio n

100

90

80

70

60

50
%

40

30

20

10

0
0 .0 0 1 0 .0 1 0 .1 1 10 100
G ra in s iz e in m m
O r ig i n a l P S D Loaded P S D O v e rfl o w P S D

T ra i lin g S u c tio n H o p p e r D r e d g e V 1 .3 , A p ri l 2 9 , 2 0 0 9 , 1 6 :1 9 :2 3
S m a ll T S H D (C :\P r o g ra m F il e s \T ra i lin g S u c tio n H o p p e r D r e d g e \T S H D \0 9 S m a ll.In p )
V e r y fin e s a n d d 5 0 = 0 .1 m m ( C :\P ro g r a m F ile s \T r a il in g S u c ti o n H o p p e r D re d g e \S a n d \S a n d .In p )
O p ti m u m p ro d u c tio n : 2 1 1 1 T D S , l o a d e d in : 4 3 .0 m in , o v e rfl o w l o s s e s : 1 5 6 4 T D S

Figure 8-48: The grain distribution curves, original, overflow losses and sediment for the Small TSHD.

C u m u la tiv e G ra in S iz e D is trib u tio n

100

90

80

70

60

50
%

40

30

20

10

0
0 .0 0 1 0 .0 1 0 .1 1 10 100
G ra in s iz e in m m
O r ig i n a l P S D Loaded P S D O v e rfl o w P S D

T ra i lin g S u c tio n H o p p e r D r e d g e V 1 .3 , A p ri l 2 9 , 2 0 0 9 , 1 5 :1 1 :2 5
M e g a T S H D ( C :\P r o g r a m F i le s \T r a ili n g S u c tio n H o p p e r D r e d g e \T S H D \0 9 M e g a .In p )
V e r y fin e s a n d d 5 0 = 0 .1 m m ( C :\P ro g r a m F ile s \T r a il in g S u c ti o n H o p p e r D re d g e \S a n d \S a n d .In p )
O p ti m u m p ro d u c tio n : 3 3 1 4 9 T D S , l o a d e d in : 1 0 7 .7 m in , o v e rfl o w l o s s e s : 2 4 3 6 8 T D S

Figure 8-49: The grain distribution curves, original, overflow losses and sediment for the Mega TSHD.

Copyright © Dr.ir. S.A. Miedema TOC Page 267 of 304


Introduction Dredging Engineering.

T h e c u m u l a t iv e e ff ic i e n c y a s a fu n c ti o n o f th e m ix t u r e c o n c e n tr a ti o n
1 .0 0 1 .0 0

0 .9 0 0 .9 0

C u m u la tiv e E ffic ie n c y
C u m u la tiv e E ffic ie n c y

0 .8 0 0 .8 0

0 .7 0 0 .7 0

0 .6 0 0 .6 0

0 .5 0 0 .5 0

0 .4 0 0 .4 0

0 .3 0 0 .3 0

0 .2 0 0 .2 0

0 .1 0 0 .1 0

0 .0 0 0 .0 0
0 .0 0 0 .0 5 0 .1 0 0 .1 5 0 .2 0 0 .2 5 0 .3 0 0 .3 5 0 .4 0 0 .4 5 0 .5 0

M i x tu r e c o n c e n tr a ti o n
1 0 0 % fille d w it h w a te r 5 0 % f ille d w ith w a t e r 0 % fille d w ith w a te r

T h e c u m u l a t iv e o v e r f l o w lo s s e s a s a fu n c ti o n o f th e m ix t u r e c o n c e n tr a ti o n
1 .0 0 1 .0 0

C u m u la tiv e o v e r flo w
lo s s e s

0 .9 0 0 .9 0

0 .8 0 0 .8 0

0 .7 0 0 .7 0
C u m u la tiv e o v e r flo w

0 .6 0 0 .6 0

0 .5 0 0 .5 0

0 .4 0 0 .4 0

0 .3 0 0 .3 0
lo s s e s

0 .2 0 0 .2 0

0 .1 0 0 .1 0

0 .0 0 0 .0 0
0 .0 0 0 .0 5 0 .1 0 0 .1 5 0 .2 0 0 .2 5 0 .3 0 0 .3 5 0 .4 0 0 .4 5 0 .5 0

M i x tu r e c o n c e n tr a ti o n
1 0 0 % fille d w it h w a te r 5 0 % f ille d w ith w a t e r 0 % fille d w ith w a te r

T ra i lin g S u c tio n H o p p e r D r e d g e V 1 .3 , A p r il 2 9 , 2 0 0 9 , 1 6 :2 0 :5 4
S m a ll T S H D (C :\P r o g ra m F il e s \T r a ili n g S u c tio n H o p p e r D r e d g e \T S H D \0 9 S m a ll .In p )
V e r y fin e s a n d d 5 0 = 0 .1 m m (C :\P r o g ra m F il e s \T ra i lin g S u c tio n H o p p e r D r e d g e \S a n d \S
O p ti m u m p ro d u c tio n : 2 1 1 1 T D S , lo a d e d in : 4 3 .0 m i n , o v e r flo w lo s s e s : 1 5 6 4 T D S

Figure 8-50: The overflow losses compared with an analytical model for the Small TSHD.

Page 268 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

T h e c u m u l a t iv e e ff ic i e n c y a s a fu n c ti o n o f th e m ix t u r e c o n c e n tr a ti o n
1 .0 0 1 .0 0

0 .9 0 0 .9 0

C u m u la tiv e E ffic ie n c y
C u m u la tiv e E ffic ie n c y

0 .8 0 0 .8 0

0 .7 0 0 .7 0

0 .6 0 0 .6 0

0 .5 0 0 .5 0

0 .4 0 0 .4 0

0 .3 0 0 .3 0

0 .2 0 0 .2 0

0 .1 0 0 .1 0

0 .0 0 0 .0 0
0 .0 0 0 .0 5 0 .1 0 0 .1 5 0 .2 0 0 .2 5 0 .3 0 0 .3 5 0 .4 0 0 .4 5 0 .5 0

M i x tu r e c o n c e n tr a ti o n
1 0 0 % fille d w it h w a te r 5 0 % f ille d w ith w a t e r 0 % fille d w ith w a te r

T h e c u m u l a t iv e o v e r f l o w lo s s e s a s a fu n c ti o n o f th e m ix t u r e c o n c e n tr a ti o n
1 .0 0 1 .0 0

C u m u la tiv e o v e r flo w
lo s s e s

0 .9 0 0 .9 0

0 .8 0 0 .8 0

0 .7 0 0 .7 0
C u m u la tiv e o v e r flo w

0 .6 0 0 .6 0

0 .5 0 0 .5 0

0 .4 0 0 .4 0

0 .3 0 0 .3 0
lo s s e s

0 .2 0 0 .2 0

0 .1 0 0 .1 0

0 .0 0 0 .0 0
0 .0 0 0 .0 5 0 .1 0 0 .1 5 0 .2 0 0 .2 5 0 .3 0 0 .3 5 0 .4 0 0 .4 5 0 .5 0

M i x tu r e c o n c e n tr a ti o n
1 0 0 % fille d w it h w a te r 5 0 % f ille d w ith w a t e r 0 % fille d w ith w a te r

T ra i lin g S u c tio n H o p p e r D r e d g e V 1 .3 , A p r il 2 9 , 2 0 0 9 , 1 5 :1 2 :2 1
M e g a T S H D ( C :\P r o g r a m F i le s \T r a il in g S u c ti o n H o p p e r D re d g e \T S H D \0 9 M e g a .In p )
V e r y fin e s a n d d 5 0 = 0 .1 m m (C :\P r o g ra m F il e s \T ra i lin g S u c tio n H o p p e r D r e d g e \S a n d \S
O p ti m u m p ro d u c tio n : 3 3 1 4 9 T D S , lo a d e d in : 1 0 7 .7 m i n , o v e r flo w lo s s e s : 2 4 3 6 8 T D S

Figure 8-51: The overflow losses compared with an analytical model for the Mega TSHD.

Copyright © Dr.ir. S.A. Miedema TOC Page 269 of 304


Introduction Dredging Engineering.

8.11. Waterjets in the Draghead.


Instead of cutting sand with a draghead, it is also possible to fluidize the sand with waterjets. The fluidization
process depends on the differential jet pressure and the jet flow. Where the jet flow can be determined by using
the Bernouilli equation applied to the differential jet pressure and the internal cross section of the nozzle. The
momentum of the flow determines the penetration depth of the water jet. If the jets are close enough and the
penetration depth matches the cutting depth, the cutting process is replaced by a jetting process. Since the total
efficiency of the jetting process is much higher than the total efficiency of the cutting process and the required net
jet power is also much less than the net cutting power, the installed power for water jet pumps is much less than
the power used by the propulsion system to generate the cutting forces, resulting in a much smaller fuel
consumption.

Waterjets however have the disadvantage that the penetration depth depends on the type of sand, especially on the
permeability of the sand. A smaller value of the permeability will reduce the penetration depth. If the penetration
depth is smaller than the layer thickness that would be cut by the blades in the draghead, the blades will still have
to cut the difference. If the permeability is larger its possible that more sand is fluidized than the amount entering
the draghead based on the depth of the tip of the blades.

A model is presented here based on the assumption that the energy required to fluidize the sand is equal to the
energy required to cut the sand at zero waterdepth with a small blade angle. Since the working principle of waterjets
is based on the differential pressure over the nozzle of a waterjet, this does not depend on the waterdepth. The
model assumes all the sand up to the penetration depth is fluidized, where in reality this may not be the case.

The Model:
The situ production Qs of 1 draghead equals the layer thickness (the penetration depth of the jets) hi times the
width of the blades w times the trailing speed vc.

Qs  hi  w  vc (8-75)

So the penetration depth equals:

Qs
hi  (8-76)
w  vc

The situ production of 1 draghead also equals the installed jetpump power Pj divided by the specific energy Esp:

Pj
Qs  (8-77)
E sp

This gives for the penetration depth:

Pj 1
hi   (8-78)
E sp w  vc

The flow Qj leaving the waterjet equals the number of jets nj times the jet water velocity vj times the cross section
of the jet opening, including the effect of the contraction coefficient α:

 2
Q j  nj vj 
4

  Dj  (8-79)

According to Bernouilli the jet water velocity vj can be expressed in terms of the pressure difference over the
nozzle Δpj, so:

1/2
 2  pj 
vj    (8-80)
 l 
 

Page 270 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

So the total jet water flow is:

1/2
 2  p j   2
Q j  nj 
 l



4

  D j  (8-81)
 
The jet water power equals the pressure Δpj times the flow Qj giving:

1/2
 2  pj   2
Pj   p j  Q j  pj nj 
 l



4

  D j  (8-82)
 

Now the situ production of 1 draghead Qs is:

1/2
 2  p j   2
p j n j 
 l
 
4

  Dj  (8-83)
 
Qs 
E sp

The penetration depth hi is:

1/2
 2  p j   2
p j n j 
 l
 
4

  Dj  (8-84)
 
hi 
E sp  w  v c

The question is now, which specific energy Esp to use in this equation. Assuming the specific energy for jetting
equals the specific energy for non-cavitating cutting, this gives:

l  g hi  vc  
E sp  c1  (8-85)
km

This gives for the penetration depth hi:

1/2
 2  p j   2
pj nj 
 l
 
4

  Dj  (8-86)
  km
hi  
c1   l  g  h i 
2
vc w 

Or:

1/2
 2  p j   2
pj nj 
 l
 
4

  Dj  (8-87)
2   km
hi  
c1   l  g  v c  w
2 

There are still a number of unknowns in this equation. The contraction coefficient α has a value close to 0.85. The
density of the water ρl has a value of about 1.025 ton/m3. The ratio of the width of the draghead w to the number
of jets nj is about 0.2 m. Since the jetting process does not know about blade angles, a c1 value has to be chosen
calibrated on experiments. A c1 value of 0.12 is found, matching a 15º blade.

The ratio of the mean permeability km to the dilatation ε required some investigation. The mean permeability is
the average of the initial (situ) permeability ki and the maximum permeability kmax after the sand has passed the
shear plane in sand cutting. The dilatation is the volume increase due to the shearing of the sand. The mean
permeability and the dilatation are related. An increase in dilatation also results in an increase of the maximum

Copyright © Dr.ir. S.A. Miedema TOC Page 271 of 304


Introduction Dredging Engineering.

permeability and thus an increase of the mean permeability. Using the Kozeny Carman equation to determine the
permeability, the following is found:

km
 10  k i (8-88)

Giving for the penetration depth:

1/2
 2  pj   2
p j n j 
 l
 
4

 Dj  (8-89)
2  
hi   10  k i
2
c1   l  g  vc w

This can be simplified to:

1/2
 2   2
    3/2 2 3/2 2
 l  4 pj D j ki pj D j ki
2
hi  10    3 2 .8 5  (8-90)
w 2 2
c1   l  g  vc vc
nj

The penetration depth is now:

3/4 1/2
pj D j ki
h i  5 .7 3  (8-91)
vc

The situ prodiction can be determined by:

3/4 1/2
Q s  5 .7 3   p j Dj ki w (8-92)

So the situ production when using waterjets does not depend on the trailing speed vc. This makes sense, since the
jet power is a constant.

The jet power required is now:

1/2 1/2
 2   2 3/2 2  2   2 3/2 2 w
Pj       p j D j nj       pj D j  (8-93)
 l  4  l  4 0 .2

Giving, using the different values:

3/2 2
Pj  4   p j Dj w (8-94)

Since the jetpumps also have an efficiency (0.8-0.9) the installed jetpump power for 1 draghead is:

3/2 2
4  pj D j w
Pj  (8-95)
j

The model is calibrated for a differential pressure of 6 bar, a nozzle diameter of 3 cm, a permeability of 0.0001
m/s giving a 0.2 m penetration depth and requiring about 13 kW power per nozzle. For a draghead of 3 m width
(suction pipe diameter about 1 m) and 15 nozzles, the required power is about 200 kW and the situ production is
about 0.6 m3/s per draghead. A differential pressure of 10 bar and a nozzle diameter of 5 cm would require about
80 kW power per nozzle (1200 kW total), giving a situ production of 1.5 m3/s. For the case considered this would
give a mixture density of about 1.3 ton/m3.

Page 272 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

8.12. Conclusions & Discussion.


The Camp and Dobbins model can be used to estimate loading time and overflow losses; however, the model
should be tuned with measurements of the overflow rate in tons/sec as well as the particle size distribution in the
overflow, as a function of time. The model can then also be used for the calculation of the decaying of the overflow
plume in the dredging area.
If the model is used for the calculation of the production rate of the dredge a distinction has to be made whether
the production is expressed in T.D.S./sec or in m3/sec. In the first case the theory can be applied directly, while in
the second case it has to be realized, that the overflow losses in T.D.S./sec do not always result in the same overflow
loss in m3/sec, since fine particles may situate in the voids of the bigger ones. The loss of fines does not reduce the
total volume, but increases the void ratio. Although the fines leave the hopper in this case, they do not result in a
reduction of the volume of the settled grains.
Those fractions which can be considered to apply to the overflow losses and those which do not, can be estimated
from the difference between the real particle size distribution and the optimal particle size distribution, giving a
maximum dry density, the so called Fuller distribution. If the gradient of the distribution curve for the fines is less
steep then the corresponding gradient of the Fuller distribution, than that fraction of fines will not effectively
contribute to the overflow losses if they are expressed in m3/sec. In such a case, in-situ, the fines were situated in
the voids of the courser grains. If the gradient is however steeper, the fines also form the grain matrix and the
volume of settled grains will decrease if the fines leave the hopper through the overflow.
In the model a number of assumptions are made. Except from numerical values for the parameters involved, the
Camp and Dobbins approach is used for the influence of turbulence, while separately the influence of scour is used
instead of using it as a boundary condition.

The models of Miedema & Vlasblom (1996) and van Rhee (2002C) give the same magnitude for the overflow
losses, but the shapes of the curves are different due to the differences in the physical modeling of the processes.
Due to the lower losses the computed optimal loading time will be shorter for the Vlasblom /Miedema approach.
The strong point of the van Rhee model is the accurate physical modeling, giving the possibility to model the
geometry of the hopper in great detail, but also describing the physical processes in more detail. The van Rhee
model is verified and validated with model and prototype tests and can be considered a reference model for other
models. The strong point of the Miedema/Vlasblom model is the simplicity, giving a transparent model where
result and cause are easily related.

One question before this research started, was how do the cumulative overflow losses behave when TSHD’s are
scaled from small to very big. The second question was, are that scale laws that should be applied when scaling
TSHD’s in order to create similar or maybe even identical processes.

First the answer on the second question, there are scale laws that should be applied and the main law is, to keep
the hopper load parameter constant and from there derive the scale laws for the flow and other dimensions, but
don’t scale the sand. If the scale laws are applied correctly, the simulations show that scaling the TSHD has hardly
any influence on the cumulative overflow losses and the loading processes are similar.
The overflow losses however depend strongly on the position of the grain diameter with respect to the hopper load
parameter in the particle size distribution diagram. The fraction of the sand with diameters smaller than this
diameter has a very strong relation with the cumulative overflow losses. A large silt fraction will increase these
overflow losses.

Finally we have noted that the modified Hopper Load Parameter will reduce in magnitude compared with the
unmodified Hopper Load Parameter. For particles with a settling efficiency greater than 1, this will not influence
the settling efficiency, but for particles with a settling efficiency near 1 or smaller than 1, this may increase the
settling efficiency slightly. So the sedimentation velocity in this respect has a positive effect on the cumulative
settling efficiency. The current model seems to give rather accurate predictions. This conclusion is based on the
comparison with the van Rhee model on one hand and the comparison with real data on the other hand.

Four effects are considered that were not part of the original Miedema & Vlasblom (1996) model, based on the
Camp model. Those effects have been added later to the model by Miedema (2008A), (2008B), (2009A), (2009B),
(2010) and Miedema & van Rhee (2007).
 Equations (8-25) and (8-29) give a good estimate of the thickness of the layer of water above the overflow
level and Figure 8-15 proves that this estimate is accurate.
 The Shields approach is based on a fundamental force and moment equilibrium on grains and has been proven
by many scientists in literature. Now the question is, which Shields curve to use. Figure 8-52 shows 7 levels
of erosion as defined by Delft Hydraulics (1972). To decide which of these 7 levels is appropriate for the

Copyright © Dr.ir. S.A. Miedema TOC Page 273 of 304


Introduction Dredging Engineering.

physics of the final stage of hopper loading, these physics should be examined. During this final stage, a high
density mixture is flowing over the sediment. Part of the particles in this mixture flow will settle, part will not
settle because the settling velocity is to low and part will not settle because of erosion and suspension. This
process differs from the erosion process in the fact that there is not water flowing over the sediment, but a
high density mixture. In fact the mixture is already saturated with particles and it is much more difficult for a
particle to get eroded that in a clean water flow. One could call this hindered erosion. From the experience
until now with the erosion model described (Miedema & van Rhee (2007)) and comparing it with other
models, level 7 from Figure 8-52 should be chosen, this level is achieved by using β=0.475.
 The concentration of the mixture above the bed, often called the near bed concentration cb, can be estimated
with equation (8-96), and based on a black box approach. This concentration is used to determine the hindered
settling effect on the settling velocity. Although equation (8-96) will not give the near bed concentration at a
certain place at a certain time, it is derived for the entire hopper and loading cycle, it’s a good estimate for
determining the cumulative overflow losses.
 The storage, time delay or buffer effect can be implemented by using equation (8-30). Miedema & van Rhee
(2007) compared both the Miedema & Vlasblom (1996) model, including the features as discussed here, and
the sophisticated 2DV model, van Rhee (2002C). The result is shown in Figure 8-18. It is clear from this
figure that there is a difference between the two methods if the storage effect is omitted in the Miedema &
Vlasblom model, but including this storage effect gives almost the same results.
 It looks like the modified model gives results that match the van Rhee (2002C) model closely; of course the
models are compared for just a few cases, specifically regarding the grain distributions used. This is
remarkable because the physics of the two models are different. The van Rhee (2002C) model is based on the
density flow as shown in Figure 8-6 and Figure 8-7, where there is an upward flow in the hopper. The modified
model as presented here is based on the old Camp theory and assumes a uniform inflow of particles over the
height of the hopper, as shown in Figure 8-20, a horizontal flow of the mixture and vertical downward
transport of particles. So it seems that the dominating parameter in both models is the so called hopper load
parameter, since this is the upward flow velocity in the van Rhee model and it is the settling velocity of a
particle entering the hopper at the top and just reaching the sediment at the other end of the hopper in the
Miedema & Vlasblom model.

Using the equations to determine the near bed concentration as derived here are based on known cumulative
overflow losses and should thus not be used to predict overflow losses because that is a self-fulfilling prophecy.
The modeling should be used to verify experiments where the near bed concentration is measured.
The use of the sedimentation or bed rise velocity to determine the sedimentation process when loading a TSHD
with sand can only give good predictions if the correct near bed concentration is used and measured. Using the
assumption that the near bed concentration equals the inflowing mixture concentration may lead to results that do
not obey the conservation of mass principle.
Using the empirical equation (8-97) of van Rhee (2002C) to predict the overflow losses with the assumption that
cb=cin is a good first approximation, but with some restrictions. It should be noted that van Rhee used the
assumption of cb=cin to find this equation by curve fitting. The dimensionless overflow rate S* in this equation has
to be considered to be the reciprocal of the settling efficiency, that is the correct physical meaning.
The analytical model derived in this paper matches this empirical equation, but has the advantage that sands with
different grading can be taken into account.
The model derived for the sedimentation velocity, the near bed concentration and the overflow losses matches both
the experiments as carried out by van Rhee (2002C) and Ooijens et al. (2001).
The model however is very sensitive for the values of the parameters a and b describing the PSD in equation
(8-98), but with correct values, the model gives a very good prediction of the cumulative overflow losses.

cb cb  cu m
   
(8-96)
c b ed   c in   cu m    p 

*
o v c u m  0 .3 9  ( S '  0 .4 3 ) (8-97)

lo g ( d )  a  p  b (8-98)

Page 274 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

8.13. Nomenclature
a Steepness of the PSD mm
b Offset of the PSD mm
b Width of the weir m
cb Near bed concentration -
cbed Bed/sediment concentration -
cin Volume concentration -
cv , ci Volumetric concentration -
Ce Dimensionless discharge (contraction) coefficient with a value near 0.6. -
Cd Coefficient -
CD Drag coefficient -
CL Lift coefficient -
d Grain diameter mm
do Grain diameter matching the hopper load parameter mm
d50 Grain diameter at 50% of PSD mm
ds Grain diameter (scour) m
FD Drag force kN
FL Lift force kN
Fw Submerged weight kN
g Gravitational constant (9.81) m/sec2
h Height m
h is the overfall height (measured about a distance of 5·h upstream from the crest) m
hmax Maximum water layer thickness m
H Height of hopper m
Hw Height of the water above the sediment m
H* Dimensionless hopper load parameter -
L Length of basin m
M Height of the weir crest above the headwater bottom m
n Porosity -
ov Overflow losses -
ovcum Cumulative overflow losses -
p Fraction of grains -
po Fraction of grains that settle partially (excluding turbulence) -
pfs , ps Fraction of grains that do no settle due to scour or fines -
p0 Atmospheric pressure kPa
Q Mixture flow m3/sec
Qin, out Mixture flow (in or out) m3/sec
Qm Mixture flow (mass) ton/sec
Rd Relative density -
R Reduction factor -
so Flow velocity in basin m/sec
ss Scour velocity m/sec
S* Dimensionless overflow rate -
S Sedimentation flux
t, T Time sec
TDS Tonnes dry solid ton
u* Shear velocity m/sec
Ucr Critical velocity above bed m/sec
v Mean velocity in the headwater this is equal to Q/b (M + h) m/sec
vc Settling velocity including hindered settling m/sec
vo Hopper load parameter m/sec
vs Settling velocity of individual particle m/sec
vsed Sedimentation/bed rise velocity m
W Width of basin m
α Fraction of hopper to be filled with mixture at start of loading process -
α Velocity factor -
β Power for hindered settling -

Copyright © Dr.ir. S.A. Miedema TOC Page 275 of 304


Introduction Dredging Engineering.

β Height factor -
ε Fraction of hopper filled with sediment when reaching the overflow -
ρf Density of fluid ton/m3
ρq Density of particles (quarts=2.65) ton/m3
ρw Density of water (1.025) ton/m3
m Density of a sand/water mixture ton/m3
q Density of quarts ton/m3
s Density of sediment ton/m3
η Settling efficiency -
ηcum Cumulative settling efficiency -
ηg Settling efficiency individual grain -
ηb Settling efficiency for basin -
ηt Turbulence settling efficiency for individual grain -
ηp Settling efficiency individual particle -
λ Concentration ratio cb/cin -
 Viscous friction coefficient -
κ Concentration ratio cin/cbed -
κ Ratio mixture concentration versus bed concentration -
μ Settling velocity factor -
μ Friction coefficient -
τ Time constant sec
ν Kinematic viscosity St
θ Shields parameter -

S h ie ld s D ia g ra m
1
S h ie ld s P a r a m e te r

0 .1

0 .0 1
0 .1 1 10 100 1000

Re*
S o u ls b y M ie d e m a B e ta = 0 .4 7 5 B e ta = 0 .5 2 5 B e ta = 0 .6 0 0 B e ta = 0 .7 0 0 B e ta = 0 .8 0 0 B e ta = 0 .9 0 0 B e ta = 1 .0 0 0

Figure 8-52: The 7 levels of erosion according to Delft Hydraulics (1972).

Page 276 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

8.14. Notes.

Copyright © Dr.ir. S.A. Miedema TOC Page 277 of 304


Introduction Dredging Engineering.

Page 278 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

Copyright © Dr.ir. S.A. Miedema TOC Page 279 of 304


Introduction Dredging Engineering.

Page 280 of 304 TOC Copyright © Dr.ir. S.A. Miedema


The Trailing Suction Hopper Dredge.

Copyright © Dr.ir. S.A. Miedema TOC Page 281 of 304


Introduction Dredging Engineering.

Page 282 of 304 TOC Copyright © Dr.ir. S.A. Miedema


References.

Chapter 9: References.
Abelev, A., & Valent, P. (2010). Strain rate dependency of strength of soft marine deposits of the Gulf of Mexico.
Stennis Space Center, MS 39529, USA.: Naval Research Laboratory.
Abulnaga, B. E. (2002). Slurry Systems Handbook. USA: McGraw Hill.
Azamathulla, H. M., & Ahmad, Z. (2013). Estimation of critical velocity for slurry transport through pipeline using
adaptive neuro-fuzzy interference system and gene-expression programming. Journal of Pipeline Systems
Engineering and Practice., 131-137.
Babcock, H. A. (1970). The sliding bed flow regime. Hydrotransport 1 (pp. H1-1 - H1-16). Bedford, England:
BHRA.
Bagnold, R. A. (1954). Experiments on a gravity free dispersion of large solid spheres in a Newtonian fluid under
shear. Proceedings Royal Society, Vol. A225., 49-63.
Bagnold, R. A. (1957). The flow of cohesionless grains in fluids. Phil. Trans. Royal Society, Vol. A249, 235-297.
Bain, A. G., & Bonnington, S. T. (1970). The hydraulic transport of solids by pipeline. Pergamon Press.
Barker, A., Sayed, M., & Carrieres, M. (2004). Determination of iceberg draft, mass and cross sectional areas.
14th international offshore and polar engineering conference. Toulon, France.
Barrette, P. (2011). Offshore pipeline protection against seabed scouring, an overview. Cold Regions Science
Technology, Vol. 69., 3-20.
Becker, S., Miedema, S., Jong, P., & Wittekoek, S. (1992). On the Closing Process of Clamshell Dredges in Water
Saturated Sand. WODCON XIII (p. 22 pages). Bombay, India: WODA.
Becker, S., Miedema, S., Jong, P., & Wittekoek, S. (1992, September). The Closing Process of Clamshell Dredges
in Water Saturated Sand. Terra et Aqua, No. 49, 22 pages.
Berg, C. H. (1998). Pipelines as Transportation Systems. Kinderdijk, the Netherlands: European Mining Course
Proceedings, IHC-MTI.
Berg, C. v. (2013). IHC Merwede Handbook for Centrifugal Pumps & Slurry Transportation. Kinderdijk,
Netherlands: IHC Merwede.
Berman, V. P. (1994). Gidro i aerodinamiceskie osnovy rascota truboprovodnych sistem gidrokontejnernogo i
vysokonapornogo pnevmaticeskogo transporta. Lugansk: East Ukrainian State University.
Biot, M. (1941). General theory of three dimensional consolidation. Journal of Applied Physics, vol. 12., 155-164.
Bishop, A. (1966). The strength of soils as engineering materials. Geotechnique, vol. 16, no. 2., 91-128.
Bisschop, F., Miedema, S. A., Rhee, C. v., & Visser, P. J. (2014). Erosion experiments on sand at high velocities.
To be submitted to the Journal of Hydraulic Engineering, 28.
Blasco, S. M., Shearer, J. M., & Myers, R. (1998). Seabed scouring by sea ice: scouring process and impact rates.
1st ice scour and arctic marine pipelines workshop, 13th international symposium on Okhotsk Sea and
Sea Ice., (pp. 53-58). Mombetsu, Hokkaido, Japan.
Blatch, N. S. (1906). Discussion of Works for the purification of the water supply of Washington D.C.
Transactions ASCE 57., 400-409.
Blythe, C., & Czarnotta, Z. (1995). Determination of hydraulic gradient for sand slurries. 8th International Freight
Pipeline Society Symposium, (pp. 125-130). Pittsburg, USA.
Bonneville, R. (1963). essais de synthese des lois debut d'entrainment des sediment sous l'action d'un courant en
regime uniform. Chatou: Bulletin Du CREC, No. 5.
Bonnington, S. T. (1961). Estimation of Pipe Friction Involved in Pumping Solid Material. BHRA, TN 708
(December 1961).
Boothroyde, J., Jacobs, B. E., & Jenkins, P. (1979). Coarse particle hydraulic transport. Hydrotransport 6: 6th
International Conference on the Hydraulic Transport of Solids in Pipes. (p. Paper E1). BHRA.
Bourbonnair, J., & Lananyi, B. (1985). The mechanical behaviour of a frozen clay down to cryogenic temperatures.
4th symposium on ground freezing., (pp. 237-244). Sapporo, Japan.
Brakel, J. (1981). Mathematisch model voor de krachten op een roterende snijkop van een in zeegang werkende
snijkopzuiger. Delft, Netherlands: Delft University of Technology - ScO/80/96.
Brauer, H. (1971). Grundlagen der einphasen- und mehrphasenstromungen. Verslag Sauerlander.
Bree, S. E. (1977). Centrifugal DredgePumps. Ports & Dredging (10 issues, IHC Holland).
Brooks, F. A., & Berggren, W. (1944). Remarks on turbulent transfer across planes of zero momentum exchange.
Transactions of the American Geophysics Union, Pt. VI., 889-896.
Brownlie, W. (1981). Compilation of alluvial channel data: laboratory and field, Technical Report KH-R-43B.
Pasadena, California, USA: California Institute of Technology.
Buffington, J. M. (1999). The legend of A.F. Shields. Journal of Hydraulic Engineering, 125, 376–387.
Buffington, J. M., & Montgomery, D. R. (1997). A systematic analysis of eight decades of incipient motion studies,
with special reference to gravel-bedded rivers. Water Resources Research, 33, 1993-2029.
Butterfield, R., & Andrawes, K. (1972). On the angles of friction between sand and plane surfaces. Journal of
Terramechanics, vol. 8, no. 4., 15-23.

Copyright © Dr.ir. S.A. Miedema TOC Page 283 of 304


Introduction Dredging Engineering.

Camenen, B., & Larson, M. (2013). Accuracy of Equivalent Roughness Height Formulas in Practical Applications.
Journal of Hydraulic Engineering., 331-335.
Camenen, B., Bayram, A. M., & Larson, M. (2006). Equivalent roughness height for plane bed under steady flow.
Journal of Hydraulic Engineering, 1146-1158.
Camp, T. (1936). Study of rational design of settling tanks. Sewage Works Journal 8-5., 742-758.
Camp, T. (1946). Sedimentation and the design of settling tanks. ASCE Transactions, 895-936.
Camp, T. (1953). Studies of sedimentation design. Sewage & Industrial Wastes 25, 1-14.
Carman, P. (1937). Fluid flow through granular beds. Transactions Institute Chemical Engineering, 15, 150.
Carman, P. (1956). Flow of gases through porous media. London: Butterworths Scientific Publications.
Charles, M. E. (1970). Transport of solids by pipeline. Hydrotransport 1. Cranfield: BHRA.
Chaskelberg, K., & Karlin. (1976). Rascot gidrotransporta pesanych materialov. Moskov: Gidromechanizacija.
Chen, X., & Miedema, S. A. (2014). NUMERICAL METHODS FOR MODELING THE ROCK CUTTING
PROCESS IN DEEP SEA MINING. OMAE 2014 (p. 10). San Francisco, USA: ASME.
Chin, C. O., & Chiew, Y. M. (1993). Effect of bed surface structure on spherical particle stability. Journal of
Waterway, Port, Coastal and Ocean Engineering, 119(3), 231–242.
Clift, R., Wilson, K. C., Addie, G. R., & Carstens, M. R. (1982). A mechanistically based method for scaling
pipeline tests for settling slurries. Hydrotransport 8 (pp. 91-101). Cranfield, UK.: BHRA Fluid
Engineering.
Clift, R., Wilson, K., Addie, G., & Carstens, M. (1982). A mechanistically based method for scaling pipeline tests
for settling slurries. Hydrotransport 8 (pp. 91-101). Cranfield, UK.: BHRA.
Colebrook, C. F., & White, C. M. (1937). Experiments with Fluid Friction in Roughened Pipes. Proceedings of
the Royal Society of London. Series A, Mathematical and Physical Sciences 161 (906). (pp. 367-381).
London: Royal Society of London.
Coleman, N. L. (1967). A theoretical and experimental study of drag and lift forces acting on a sphere resting on
a hypothetical stream bed. International Association for Hydraulic Research,12th Congress, 3, pp. 185-
192.
Condolios, E., & Chapus, E. E. (1963A). Transporting Solid Materials in Pipelines, Part I. Journal of Chemical
Engineering, Vol. 70(13)., 93-98.
Condolios, E., & Chapus, E. E. (1963B). Designing Solids Handling Pipelines Part II. Journal of Chemical
Engineering, Vol. 70(14)., 131-138.
Condolios, E., & Chapus, E. E. (1963C). Operating solids pipelines, Part III. Journal of Chemical Engineering,
Vol. 70(15)., 145-150.
Coulomb, C. (1776). Essai sur une application des regles des maximis et minimis a quelques problemes de statique
relatifs a l'architecture. Academie royale des sciences, Paris, Memoires de mathematique et de physique,
vol. 7., 343-382.
Cox, C. M., Eygenraam, J. A., Granneman, C. C., & Njoo, M. (1995). A Training Si,ulator for Cutter Suction
Dredgers: Bridging the Gap between Theory and Practice. WODCON XIV (p. 10). Amsterdam, The
Netherlands.: WODA.
Crowe, C. T. (2006). MultiPhase Flow Handbook. Boca Raton, Florida, USA: Taylor & Francis Group.
Davies, J. T. (1987). Calculation of critical velocities to maintain solids in suspension in horizontal pipes. Chemical
Engineering Science, Vol. 42(7)., 1667-1670.
Detournay, E., & Atkinson, C. (2000). Influence of pore pressure on the drilling response in low permeability shear
dilatant rocks. International Journal of Rock Mechanics & Mining Sciences, vol. 37., 1091-1101.
Dey, S. (1999). Sediment threshold. Applied Mathematical Modelling, 399-417.
Dey, S. (2003). Incipient motion of bivalve shells on sand beds under flowing water. Journal of Hydraulic
Engineering, 232-240.
Dey, S. (2014). Fluvial Hydrodynamics. Kharagpur, India: Springer, GeoPlanet: Earth and Planetary Sciences.
DHL. (1972). Systematic Investigation of Two Dimensional and Three Dimensional Scour, Report M648/M863.
Delft, Netherlands: Delft Hydraulics Laboratory.
Di Filice, R. (1999). The sedimentation velocity of dilute suspensions of nearly monosized spheres. International
Journal of Multiphase Flows 25, 559-574.
Dobbins, W. (1944). Effect of Turbulence on Sedimentation. ASCE Transactions Vol. 109, No. 2218, 629-656.
Doron, P., & Barnea, D. (1993). A three layer model for solid liquid flow in horizontal pipes. International Journal
of Multiphase Flow, Vol. 19, No.6., 1029-1043.
Doron, P., & Barnea, D. (1995). Pressure drop and limit deposit velocity for solid liquid flow in pipes. Chemical
Engineering Science, Vol. 50, No. 10., 1595-1604.
Doron, P., & Barnea, D. (1996). Flow pattern maps for solid liquid flow in pipes. International Journal of
Multiphase Flow, Vol. 22, No. 2., 273-283.
Doron, P., Granica, D., & Barnea, D. (1987). Slurry flow in horizontal pipes, experimental and modeling.
International Journal of Multiphase Flow, Vol. 13, No. 4., 535-547.

Page 284 of 304 TOC Copyright © Dr.ir. S.A. Miedema


References.

Doron, P., Simkhis, M., & Barnea, D. (1997). Flow of solid liquid mixtures in inclined pipes. International Journal
of Multiphase Flow, Vol. 23, No. 2., 313-323.
Duckworth, & Argyros. (1972). Influence of density ratio on the pressure gradient in pipes conveying suspensions
of solids in liquids. Hydrotransport 2. Coventry: BHRA.
Durand, R. (1953). Basic Relationships of the Transportation of Solids in Pipes - Experimental Research.
Proceedings of the International Association of Hydraulic Research. Minneapolis.
Durand, R., & Condolios, E. (1952). Etude experimentale du refoulement des materieaux en conduites en
particulier des produits de dragage et des schlamms. Deuxiemes Journees de l'Hydraulique., 27-55.
Durand, R., & Condolios, E. (1952). Etude experimentale du refoulement des materieaux en conduites en
particulier des produits de dragage et des schlamms. (Experimental study of the discharge pipes
materieaux especially products of dredging and slurries). Deuxiemes Journees de l'Hydraulique., 27-55.
Durand, R., & Condolios, E. (1956). Donnees techniques sur le refoulement des mixture en conduites. Revue de
lÍndustriele Minerale, no. 22F, 460-481.
Durand, R., & Condolios, E. (1956). Technical data on hydraulic transport of solid materials in conduits. Revue
de L'Industrie Minerale, Numero Special 1F.
Durepaire, M. P. (1939). Contribution a létude du dragage et du refoulement des deblais a état de mixtures. Annales
des ponts et chaussees, Memoires I., 165-254.
Egiazarof, I. (1965). Calculation of non-uniform sediment concentrations. Journal of the Hydraulic Division,
ASCE, 91(HY4), 225-247.
Einstein, A. (1905). On the motion of small particles suspended in liquids at rest required by the molecular kinetic
theory of heat. Annalen der Physik Vol.17., 549-560.
Ellis, H. S., & Round, G. F. (1963). Laboratory studies on the flow of Nickel-Water suspensions. Canadian Journal
on Minerals & Mettalurgy, Bull. 56.
Engelund, F., & Hansen, E. (1967). A monograph on sediment transport to alluvial streams. Copenhagen: Technik
Vorlag.
Evans, I. (1962). A theory on basic mechanics of coal ploughing. International symposium on mining research.,
(pp. 761-798).
Evans, I. (1964). The force required to cut coal with blunt wedges. Mining Research Establishment Isleworth.
Evans, I. (1965). The force required to cut coal with blunt wedges. International journal of rock mechanics and
mining science., 1-12.
Evans, I., & Pomeroy, C. (1966). The strength, fracture and workability of coal. Pergamon Press.
Fairhurst, C. (1964). On the validity of the Brazilian test for brittle materials. International Journal of Rock
Mechanics & Mining Sciences, vol. 1., 535-546.
Fenton, J. D., & Abbott, J. E. (1977). Initial movement of grains on a stream bed: The effect of relative protrusion.
Proceedings of Royal Society, 352(A), pp. 523–537. London.
Fowkes, R. S., & Wancheck, G. A. (1969). Materials handling research: Hydraulic transportation of coarse
solids. U.S. Department of the interior, Bureau of Mines, Report 7283.
Franzi, G. (1941). Sul moto dei liquidi con materie solide in suspensione. Milano, Italy: Instituto di idraulica e
costrusioni idrauliche dei politechnico di Milano, No. 47.
Fuhrboter, A. (1961). Über die Förderung von Sand-Wasser-Gemischen in Rohrleitungen. Mitteilungen des
Franzius-Instituts, H. 19.
Fuhrboter, A. (1961). Über die Förderung von Sand-Wasser-Gemischen in Rohrleitungen. (On the advances of
sand -water mixtures in pipelines). Mitteilungen des Franzius-Instituts, H. 19.
Gandhi, R. (2015, February). Personal communication.
Garcia, M. H. (2008). Sedimentation Engineering (Vol. 110). ASCE Manuals & Reports on Engineering Practise
No. 110.
Garside, J., & Al-Dibouni, M. (1977). Velocity-Voidage Relationships for Fluidization and Sedimentation in
Solid-Liquid Systems. 2nd Eng. Chem. Process Des. Dev., 16, 206.
Gehking, K. (1987). Rock Testing Procedures at VA’s Geotechnical Laboratory in Zeltweg. Zeltweg, Austria.:
Voest Alphine, International Report TZU 48.
Gibert, R. (1960). Transport hydraulique et refoulement des mixtures en conduites. Annales des Ponts et
Chausees., 130(3), 307-74, 130(4), 437-94.
Gibert, R. (1960). Transport hydraulique et refoulement des mixtures en conduites. (Hydraulic transport and
discharge pipes of mixtures). Annales des Ponts et Chausees., 130(3), 307-74, 130(4), 437-94.
Gillies, D. P. (2013). Particle contributions to kinematic friction in slurry pipeline flow, MSc Thesis. University
of Alberta, Department of Chemical Engineering.
Gillies, R. G. (1993). Pipeline flow of coarse particles, PhD Thesis. Saskatoon: University of Saskatchewan.
Gillies, R. G. (2015). Personal communication.
Gillies, R. G., & Shook, C. A. (2000A). Modeling high concentration settling slurry flows. Canadian Journal of
Chemical Engineering, Vol. 78., 709-716.

Copyright © Dr.ir. S.A. Miedema TOC Page 285 of 304


Introduction Dredging Engineering.

Gillies, R. G., Schaan, J., Sumner, R. J., McKibben, M. J., & Shook, C. A. (2000B). Deposition velocities for
Newtonian slurries in turbulent flow. Canadian Journal of Chemical Engineering, Vol. 78., 704-708.
Gillies, R. G., Shook, C. A., & Wilson, K. C. (1991). An improved two layer model for horizontal slurry pipeline
flow. Canadian Journal of Chemical Engineering, Vol. 69., 173-178.
Gillies, R. G., Shook, C. A., & Xu, J. (2004). Modelling heterogeneous slurry flows at high velocities. The
Canadian Journal of Chemical Engineering, Vol. 82., 1060-1065.
Glasstone, S., Laidler, K., & Eyring, H. (1941). The theory of rate processes. New York: McGraw Hill.
Gogus, M., & Kokpinar, M. A. (1993). Determination of critical flow velocity in slurry transporting pipeline
systems. Proceeding of the 12th International Conference on Slurry Handling and Pipeline Transport.
(pp. 743-757). Bedfordshire, UK.: British Hydraulic Research Group.
Govier, G. W., & Aziz, K. (1972). The Flow of Complex Mixtures in Pipes. New York: University of Calgary,
Alberta, Canada.
Grace, J. (1986). Contacting modes and behaviour classification of gas-solid and other two-phase suspensions.
Canadian Journal of Chemical Engineering, vol. 64., 353-363.
Graf, W. H., & Pazis, G. C. (1977). Les phenomenes de deposition et d’erosion dans un canal alluvionnaire.
Journal of Hydraulic Research, 15, 151-165.
Graf, W. H., Robinson, M. P., & Yucel, O. (1970). Critical velocity for solid liquid mixtures. Bethlehem,
Pensylvania, USA.: Fritz Laboratory Reports, Paper 386. Lehigh University.
Graf, W. H., Robinson, M., & Yucel, O. (1970). The critical deposit velocity for solid-liquid mixtures.
Hydrotransport 1 (pp. H5-77-H5-88). Cranfield, UK: BHRA.
Grant, W. D., & Madsen, O. S. (1982). Movable bed roughness in unsteady oscillatory flow. Journal Geophysics
Resources, 469-481.
Groot, J. (1981). Rapport Beunbezinking (in Dutch). Papendrecht: Royal Boskalis Westminster.
Grunsven, F. v. (2012). Measuring the slip factor for various slurry flows using temperature calibrated Electrical
Resistance Tomography. Delft, The Netherlands.: Delft University of Technology.
Hansen, B. (1958). Line ruptures regarded as narrow rupture zones. Earth Pressure Problems, (pp. 39-48).
Brussels.
Harada, E., Kuriyama, M., & Konno, H. (1989). Heat transfer with a solid liquid suspension flowing through a
horizontal rectangular duct. Heat Transfer Jap. Res. Vol. 18., 79-94.
Hatamura, Y., & Chijiiwa, K. (1975). Analyses of the mechanism of soil cutting, 1st report. Bulletin of JSME, vol.
18, no. 120, 619-626.
Hatamura, Y., & Chijiiwa, K. (1976A). Analyses of the mechanism of soil cutting, 2nd report. Bulletin of the
JSME, vol. 19, no. 131., 555-563.
Hatamura, Y., & Chijiiwa, K. (1976B). Analyses of the mechanism of soil cutting, 3rd report. Bulletin of the JSME,
vol. 19, no. 139., 1376-1384.
Hatamura, Y., & Chijiiwa, K. (1977A). Analyses of the mechanism of soil cutting, 4th report. Bulletin of the JSME,
vol. 20, no. 139., 130-137.
Hatamura, Y., & Chijiiwa, K. (1977B). Analyses of the mechanism of soil cutting, 5th report. Bulletin of the JSME,
vol. 20, no. 141., 388-395.
Hazen, A. (1982). Some physical properties of sands and gravels with special reference to their use in filtration.
24th Annual Report, Massachusetts State Board of Health, Pub. Doc. No. 34., 539-556.
He, J., & Vlasblom, W. (1998). Modelling of saturated sand cutting with large rake angles. WODCON XV. Las
Vegas, USA: WODA.
He, J., Miedema, S., & Vlasblom, W. (2005). FEM Analyses Of Cutting Of Anisotropic Densely Compacted and
Saturated Sand. WEDAXXV/TAMU37. New Orleans, Louisiana, USA: WEDA/TAMU.
Helmons, R. L., Miedema, S. A., & Rhee, C. v. (2014). A NEW APPROACH TO MODEL HYPERBARIC ROCK
CUTTING PROCESSES. OMAE 2014 (p. 9). San Francisco, USA: ASME.
Hepy, F. M., Ahmad, Z., & Kansal, M. L. (2008). Critical velocity for slurry transport through pipeline. Dam
Engineering, Vol. XIX(3)., 169-184.
Hettiaratchi, D. (1967A). The mechanics of soil cultivation. AES, no. 3/245/C/28.
Hettiaratchi, D., & Reece, A. (1967B). Symmetrical three-dimensional soil failure. Journal of Terramechanics 4
(3)., 45-67.
Hettiaratchi, D., & Reece, A. (1974). The calculation of passive soil resistance. Geotechnique 24, no. 3., 289-310.
Hettiaratchi, D., & Reece, A. (1975). Boundary wedges in two dimensional passive soil failure. Geotechnique 25,
no. 2., 197-220.
Hettiaratchi, D., Witney, B., & Reece, A. (1966). The calculation of passive pressure in two dimensional soil
failure. Journal of Agriculture Engineering Resources 11 (2), 89-107.
Hjulstrøm, F. (1935). Studies of the morphological activity of rivers as illustrated by the River Fyris. Bulletin of
the Geological Institute, 25, 221–527. University of Uppsala.

Page 286 of 304 TOC Copyright © Dr.ir. S.A. Miedema


References.

Hjulstrøm, F. (1939). Transportation of debris by moving water, in Trask, P.D., ed., Recent Marine Sediments. A
Symposium: Tulsa, Oklahoma, American Association of Petroleum Geologists, (pp. 5-31). Tulsa,
Oklahoma.
Hoek, E., & Brown, E. T. (1988). The Hoek-Brown Failure Criterion - a 1988 Update. 15th Canadian Rock
Mechanics Symposium., (pp. 31-38).
Howard, G. W. (1938). Transportation of Sand and Gravel in a 4 Inch Pipe. Transactions ASCE Vol. 104., No.
2039., 1334-1348.
Howard, G. W. (1939). Discussion on: Transportation of sand and gravel in a four inch pipe. Transactions ASCE
Vol. 104., 157, 316, 460, 1011.
Hsu. (1986). Flow of non-colloidal slurries in pipeline. PhD Thesis, University of Illinois.
Huisman, L. (1973-1995). Sedimentation & Flotation 1973-1995. Delft, Netherlands: Delft University of
Technology.
Huisman, L. (1980). Theory of settling tanks. Delft, Netherlands: Delft University of Technology.
Hunt, J. N. (1954). The turbulent transport of suspended sediment in open channels. Royal Society of London,
Proc. Series A, Vol. 224(1158)., 322-335.
Ikeda, S. (1982). Incipient motion of sand particles on side slopes. Journal of the Hydraulic division, ASCE,
108(No. HY1).
Ismail, H. M. (1952). Turbulent transfer mechanism and suspended sediment in closed channels. Transactions of
ASCE, Vol. 117., 409-446.
Iwagaki, Y. (1956). Fundamental study on critical tractive force. Transactions of the Japanese Society of Civil
Engineers, Vol. 41, 1-21.
Joanknecht, L. (1973). Mechanisch graaf onderzoek onder water. Delft, Netherlands: Delft University of
Technology.
Joanknecht, L. (1974). Cutting forces in submerged soils. Delft, Netherlands: Delft University of Technology.
Josselin de Jong, G. (1976). Rowe's stress dilatancy relation based on friction. Geotechnique 26, no. 3, 527-534.
Jufin, A. P. (1965). Gidromechanizacija. Moskau.
Jufin, A. P., & Lopatin, N. A. (1966). O projekte TUiN na gidrotransport zernistych materialov po stalnym
truboprovodam. Gidrotechniceskoe Strojitelstvo, 9., 49-52.
Jukes, P., Kenny, S., Panapitiya, U., Jafri, S., & Eltaher, A. (2011). Arctic and harsh environment pipeline
trenching technologies and challenges. OTC. Houston: OTC.
Julien, P. (1995). Erosion and sedimentation. Cambridge University Press.
Kaitkay, P., & Lei, S. (2005). Experimental study of rock cutting under external hydrostatic pressure. Journal of
Materials Processing Technology, vol. 159., 206-213.
Karabelas, A. J. (1977). Vertical Distribution of Dilute Suspensins in Turbulent Pipe Flow. AIChE Journal, Vol.
23(4)., 426-434.
Karasik, U. A. (1973). Hydraulische Forderung von feinkorniqen Suspensionen (in russisch). Gidromechanika,
S.BO ff, Vol. 25. Kiew.
Kaushal, D. R. (1995). Prediction of particle distribution in the flow of multisized particulate slurries through
closed ducts and open channels. Delhi, India: I.I.T.. Department of Applied Mechanics, PhD Thesis.
Kaushal, D. R., & Tomita, Y. (2002B). Solids concentration profiles and pressure drop in pipeline flow of
multisized particulate slurries. International Journal of Multiphase Flow, Vol. 28., 1697-1717.
Kaushal, D. R., & Tomita, Y. (2002C). An improved method for predicting pressure drop along slurry pipeline.
Particulate Science and Technology: An International Journal, Vol. 20(4)., 305-324.
Kaushal, D. R., & Tomita, Y. (2003B). Comparative study of pressure drop in multisized particulate slurry flow
through pipe and rectangular duct. International Journal of Multiphase Flow, Vol. 29., 1473-1487.
Kaushal, D. R., & Tomita, Y. (2013). Prediction of concentration distribution in pipeline flow of highly
concentrated slurry. Particulate Science and Technology: An International Journal, Vol. 31(1)., 28-34.
Kaushal, D. R., Sato, K., Toyota, T., Funatsu, K., & Tomita, Y. (2005). Effect of particle size distribution on
pressure drop and concentration profile in pipeline flow of highly concentrated slurry. International
Journal of Multiphase Flow, Vol. 31., 809-823.
Kaushal, D. R., Seshadri, V., & Singh, S. N. (2002D). Prediction of concentration and particle size distribution in
the flow of multi-sized particulate slurry through rectangular duct. Applied Mathematical Modelling, Vol.
26., 941-952.
Kaushal, D. R., Seshadri, V., & Singh, S. N. (2003A). Concentration and particle size distribution in the flow of
multi-sized particulate slurry through rectangular duct. Journal of Hydrology & Hydromechanics, 114-
121.
Kaushal, D. R., Tomita, Y., & Dighade, R. R. (2002A). Concentration at the pipe bottom at deposition velocity for
transportation of commercial slurries through pipeline. Powder Technology Vol. 125., 89-101.
Kazanskij, I. (1967). Vyzkum proudeni hydrosmesi voda-pisek (untersuchung uber sans-wasser stromungen).
Mitteilungen des Franzius Instituts, Heft 33.

Copyright © Dr.ir. S.A. Miedema TOC Page 287 of 304


Introduction Dredging Engineering.

Kazanskij, I. (1972). Berechnungsverfahren fur die Forderung von Sand-Wasser Gemischen in Rohrleitungen.
Hannover: Franzius Institut, Heft 33.
Kazanskij, I. (1978). Scale-up effects in hydraulic transport theory and practice. Hydrotransport 5 (pp. B3: 47-
74). Cranfield, UK: BHRA Fluid Engineering.
Kazanskij, I. (1980). Vergleich verschiedener Rohrmaterialen in Bezug auf Verschleiss und Energieverbrauch
beim Hydrotransport in Rohrleitungen. VDI Berichte Nr. 371, pp. 51-58.
Kelessidis, V., & Maglione, R. (2008). Yield stress of water bentonite dispersions. Colloids and Surfaces A:
Physicochemical Engineering Aspects 318., 217-226.
Kelessidis, V., Tsamantaki, C., & Dalamarinis, P. (2007). Effect of pH and electrolyte on the rheology of aqueous
Wyoming bentonite dispersions. Applied Clay Science 38., 86-96.
Kesteren, W. G. (1995). Numerical simulations of crack bifurcation in the chip forming cutting process in rock. In
G. B. Karihaloo (Ed.), Fracture of brittle disordered materials: concrete, rock and ceramics. (pp. 505-
524). London, UK.: E&FN Spon.
King, R. P. (2002). Introduction to Practical Fluid Flow. University of Utah.: Butterworth Heineman.
Kokpinar, M. A., & Gogus, M. (2001). Critical velocity in slurry transport in horizontal pipelines. Journal of
Hydraulic Engineering, Vol. 127(9)., 763-771.
Koning, J. d. (1977). Constant Tonnage Loading System of Trailing Suction Hopper Dredges. International Course
on Modern Dredging (p. D6). The Hague, The Netherlands: Delft University of Technology & KIVI.
Koning, J. d., Miedema, S., & Zwartbol, A. (1983). Soil/Cutterhead Interaction under Wave Conditions. WODCON
X. Singapore: WODA.
Korzajev, M. (1964). Metod rascota parametrov gidrotransporta gruntov. Gidromechanizacija, Moskau.
Kozeny, J. (1927). Uber kapillare leitung des wassers in boden. Wien: Sitzungsber. Akad. Wiss. Wien, Math.
Naturwiss. Kl. Abt. 2a, 136, 271-306.
Kril, S. I. (1990). Nopernye vzvesenesuscie potoki (pressurised slurry flows). Kiev: Naukova Dumka.
Krivenko. (1970). Energieverlust in zwei phasen stromungen in hochkonzentrierten grobdispersionen.
Gidromechanica, Kiev.
Kumar, U., Mishra, R., Singh, S. N., & Seshadri, V. (2003). Effect of particle gradation on flow characteristics of
ash disposal pipelines. Powder Technology Vol. 132., 39-51.
Kumar, U., Singh, S. N., & Seshadri, V. (2008). Prediction of flow characteristics of bimodal slurry in horizontal
pipe flow. Particulate Science and Technology, Vol. 26., 361-379.
Kurihara, M. (1948). On the critical tractive force. Research Institute for Hydraulic Engineering, Report No. 3,
Vol. 4.
Lahiri, S. K. (2009). Study on slurry flow modelling in pipeline. Durgapur, India: National Institute of Technology,
Durgapur, India.
Lambe, T., & Whitman, R. (1979). Soil mechanics, SI version. New York: John Wiley & Sons.
Lane, E. W., & Kalinske, A. A. (1941). Engineering calculations of suspended sediment. Trans. Am. Geophysics
Union, Vol. 20(3)., 603-607.
Leussen, W. v., & Os, A. v. (1987 December). Basic research on cutting forces in saturated sand. Journal of
Geotechnical Engineering, vol. 113, no. 12., 1501-1516.
Leussen, W., & Nieuwenhuis, J. (1984). Soil mechanics aspects of dredging. Geotechnique 34, no. 3., 359-381.
Liefferink, D. M., Alvarez Grima, M., Miedema, S. A., Plat, R., & Rhee, C. v. (2014). Failure mechanism of
cutting submerged frozen clay in an arctic trenching process. OTC 2014. Houston, Texas, USA.: OTC.
Lobanov, V., & Joanknecht, L. (1980). The cutting of soil under hydrostatic pressure. WODCON IX. Vancouver,
Canada: WODA.
Longwell, P. A. (1977). Mechanics of Fluid Flow. New York: McGraw Hill.
Luckner, T. (2002). Zum Bewegungsbeginn von Sedimenten. Dissertation. Darmstadt, Germany: Technische
Universitat Darmstadt.
Ma, Y. (2001). Mathematical model analysis for the saturated sand cutting with large cutting angles in the non-
vavitation situation. Delft, Netherlands: Delft University of Technology, Report: 2001.BT.5581.
Ma, Y., Ni, F., & Miedema, S. (2006A). Calculation of the Blade Cutting Force for small Cutting Angles based
on MATLAB. The 2nd China Dredging Association International Conference & Exhibition, themed
Dredging and Sustainable Development. Guangzhou, China: CHIDA.
Ma, Y., Ni, F., & Miedema, S. (2006B). Mechanical Model of Water Saturated Sand Cutting at Blade Large
Cutting Angles. Journal of Hohai University, ISSN 1009-1130, CN 32-1591.
Madsen, O. S., Wright, L. D., Boon, J. D., & Chrisholm, T. A. (1993). Wind stress, bed roughness and sediment
suspension on the inner shelf during an extreme storm event. Continental Shelf Research 13, 1303-1324.
Mantz, P. A. (1977). Incipient transport of fine grains and flakes by fluids—Extended Shields diagram. Journal of
Hydraulic Division, ASCE, 103(6), 601-615.
Matousek, V. (1996). Solids Transportation in a Long Pipeline Connected with a Dredge. Terra et Aqua 62., 3-11.

Page 288 of 304 TOC Copyright © Dr.ir. S.A. Miedema


References.

Matousek, V. (1997). Flow Mechanism of Sand/Water Mixtures in Pipelines, PhD Thesis. Delft, Netherlands: Delft
University of Technology.
Matousek, V. (2004). Dredge Pumps & Slurry Transport, Lecture Notes. Delft: Delft University of Technology.
Matousek, V. (2007). Interaction of slurry pipe flow with a stationary bed. Journal of the South African Institute
of Mining and Metallurgy, 107(6)., 367-374.
Matousek, V. (2009). Predictive model for frictional pressure drop in settling-slurry pipe with stationar deposit.
Powder Technology, 367-374.
Matousek, V. (2011). Solids Transport Formula in Predictive Model for Pipe Flow of Slurry above Deposit.
Particulate Science and Technology, Vol. 29(1)., 89-106.
Matousek, V., & Krupicka, J. (2009). On equivalent roughness of mobile bed at high shear stress. Journal of
Hydrology & Hydromechanics, Vol. 57-3., 191-199.
Matousek, V., & Krupicka, J. (2010). Modeling of settling slurry flow around deposition limit velocity.
Hydrotransport 18 (p. 12). Rio de Janeiro, Brazil: BHR Group.
Matousek, V., & Krupicka, J. (2010). Semi empirical formulae for upper plane bed friction. Hydrotransport 18
(pp. 95-103). BHRA.
Matousek, V., & Krupicka, J. (2011). Unified model for coarse slurry flow with stationary and sliding bed. 15th
International Conference on Transport & Sedimentation of Solid Particles., (p. 8). Wroclaw, Poland.
Matousek, V., & Krupicka, J. (2011). Unified model for coarse slurry flow with stationary and sliding bed.
Transport and Sedimentation of Solid Particles, 15th, (p. 9). Wroclav, Poland.
Matousek, V., Krupicka, J., & Picek, T. (2013). Validation of transport and friction formulae for upper plane bed
by experiments in rectangular pipe. Journal of Hydrology and Hydromechanics., 120-125.
Meijer, K. (1981). Berekening van spanningen en deformaties in verzadigde grond. Delft, Netherlands: Delft
Hydraulics Laboratory, Report R914 part 1.
Meijer, K. (1985). Computation of stresses and strains in saturated soil, PhD Thesis. Delft, Netherlands: Delft
University of Technology.
Meijer, K., & Os, A. (1976). Pore pressures near a moving under water slope. Geotechnical Engineering Division
ASCE 102, no. GT4., 361-372.
Merchant, M. (1944). Basic mechanics of the metal cutting process. Journal of Applied Mechanics, vol. 11A., 168-
175.
Merchant, M. (1945A). Mechanics of metal cutting process, orthogonal cutting and a type 2 chip. Journal of
Applied Physics, vol. 16, no. 5., 267-275.
Merchant, M. (1945B). Mechanics of metal cutting, plasticity conditions in orthogonal cutting. Journal of Applied
Physics, vol. 16, no. 6., 318-324.
Meyer-Peter, E., & Muller, R. (1948). Formulas for bed load transport. 2nd Meeting of the International
Association for Hydraulic Structures Research., (pp. 39-64).
Miedema, S. (1981). The flow of dredged slurry in and out hoppers and the settlement process in hoppers. Delft,
The Netherlands: Delft University of Technology, ScO/81/105, 147 pages.
Miedema, S. (1995). Production Estimation Based on Cutting Theories for Cutting Water Saturated Sand.
WODCON IV (p. 30 pages). Amsterdam, The Netherlands: WODA.
Miedema, S. (2008A). An Analytical Approach to the Sedimentation Process in Trailing Suction Hopper Dredges.
Terra et Aqua 112, 15-25.
Miedema, S. (2008B). An analytical method to determine scour. WEDA XXVIII & Texas A&M 39. St. Louis, USA:
Western Dredging Association (WEDA).
Miedema, S. (2009A). The effect of the bed rise velocity on the sedimentation process in hopper dredges. Journal
of Dredging Engineering, Vol. 10, No. 1, 10-31.
Miedema, S. (2009B). A sensitivity analysis of the scaling of TSHD's. WEDA 29 & TAMU 40 Conference.
Phoenix, Arizona, USA: WEDA.
Miedema, S. (2010). Constructing the Shields curve, a new theoretical approach and its applications. WODCON
XIX (p. 22 pages). Beijing, September 2010: WODA.
Miedema, S. A. (1981). The soil reaction forces on a crown cutterhead on a swell compensated ladder. Delft, The
Netherlands: Delft University of Technology.
Miedema, S. A. (1982). The Interaction between Cutterhead and Soil at Sea. Dredging Day November 19th (p. 25
pages in Dutch). Delft, The Netherlands: Delft University of Technology.
Miedema, S. A. (1982). The mathematical modeling of the soil reaction forces on a cutterhead and the development
of the computer program DREDMO. Delft, The Netherlands: Delft University of Technology.
Miedema, S. A. (1984A). Mathematical Modeling of a Seagoing Cutter Suction Dredge. Delft, The Netherlands:
Delft University of Technology/KIVI.
Miedema, S. A. (1984B, October). The Cutting of Densely Compacted Sand under Water. Terra et Aqua, No. 28,
4-10.

Copyright © Dr.ir. S.A. Miedema TOC Page 289 of 304


Introduction Dredging Engineering.

Miedema, S. A. (1985A, September). Derivation of the Differential Equation for Sand Pore Pressures. Dredging
and Port Construction, 35.
Miedema, S. A. (1985B, July). Mathematical Modeling of the Cutting of Densely Compacted Sand Under Water.
Dredging and Port Construction, 22-26.
Miedema, S. A. (1986A). The Application of a Cutting Theory on a Dredging Wheel. WODCON XI (p. 14 pages).
Brighton, UK: WODA.
Miedema, S. A. (1986B, June). Underwater Soil Cutting: a Study in Continuity. Dredging and Port Construction,
47-53.
Miedema, S. A. (1987 September). The Calculation of the Cutting Forces when Cutting Water Saturated Sand,
PhD Thesis. Delft: Delft University of Technology.
Miedema, S. A. (1989). On the Cutting Forces in Saturated Sand of a Seagoing Cutter Suction Dredge. WODCON
XII (p. 27 pages). Orlando, Florida, USA: WODA.
Miedema, S. A. (1989, December). On the Cutting Forces in Saturated Sand of a Seagoing Cutter Suction Dredge.
Terra et Aqua, No. 41, 27 pages.
Miedema, S. A. (1992). New developments of cutting theories with respect to dredging, the cutting of clay.
WODCON XIII. Bombay, India: World Dredging Association (WODA).
Miedema, S. A. (1994). On the Snow-Plough Effect when Cutting Water Saturated Sand with Inclined Straight
Blades. ASCE Dredging 94 (p. 24 pages). Orlando, Florida, USA: ASCE.
Miedema, S. A. (1995). Dynamic Pump/Pipeline Behavior Windows. Software. Delft, The Netherlands: SAM-
Consult.
Miedema, S. A. (1995). Production Estimation Based on Cutting Theories for Cutting Water Saturated Sand.
WODCON IV (p. 30 pages). Amsterdam, The Netherlands: WODA.
Miedema, S. A. (1996). Modeling and Simulation of the Dynamic Behavior of a Pump/Pipeline System. 17th
Annual Meeting & Technical Conference of WEDA. (p. 10). New Orleans, USA.: WEDA.
Miedema, S. A. (1999). Considerations in Building and using Dredge Simulators. WEDA XIX & TAMU 31 (p. 10).
Louisville, Kentucky, USA: WEDA.
Miedema, S. A. (1999). Considerations on limits of dredging processes. 19th Annual Meeting &amp; Technical
Conference of the Western Dredging Association. Louisville, Kentucky, USA: WEDA/TAMU.
Miedema, S. A. (2000). The modelling of the swing winches of a cutter dredge in relation with simulators. Texas
A/M 32nd Annual Dredging Seminar. Warwick, Rhode Island, USA: WEDA/TAMU.
Miedema, S. A. (2003). The Existence of Kinematic Wedges at Large Cutting Angles. CHIDA Dredging Days.
Shanghai, China: CHIDA.
Miedema, S. A. (2004). The Cutting Mechanisms of Water Saturated Sand at Small and Large Cutting Angles.
International Conference on Coastal Infrastructure Development - Challenges in the 21st Century.
Hongkong: ICCD.
Miedema, S. A. (2005). The Cutting of Water Saturated Sand, the FINAL Solution. WEDAXXV/TAMU37. New
Orleans, Louisiana, USA: WEDA/TAMU.
Miedema, S. A. (2006A). The Cutting of Water Saturated Sand, the Solution. CEDA African Section: Dredging
Days. Tangiers, Morocco: CEDA.
Miedema, S. A. (2006B). The Cutting of Water Saturated Sand, the Solution. The 2nd China Dredging Association
International Conference & Exhibition, themed Dredging and Sustainable Development. Guangzhou,
China: CHIDA.
Miedema, S. A. (2009). New Developments Of Cutting Theories With Respect To Dredging, The Cutting Of Clay
And Rock. WEDA XXIX/Texas A/M 40. Phoenix, Arizona, USA: WEDA/TAMU.
Miedema, S. A. (2010). New Developments of Cutting Theories with respect to Offshore Applications. ISOPE (p.
8). Beijing, China.: ISOPE.
Miedema, S. A. (2012A). Constructing the Shields Curve: Part A Fundamentals of the Sliding, Rolling and Lifting
Mechanisms for the Entrainment of Particles. Journal of Dredging Engineering, Vol. 12., 1-49.
Miedema, S. A. (2012B). Constructing the Shields Curve: Part B Sensitivity Analysis, Exposure & Protrusion
Levels, Settling Velocity, Shear Stress & Friction Velocity, Erosion Flux and Laminar Main Flow.
Journal of Dredging Engineering, Vol. 12., 50-92.
Miedema, S. A. (2013). An overview of theories describing head losses in slurry transport. A tribute to some of
the early researchers. OMAE 2013, 32nd International Conference on Ocean, Offshore and Arctic
Engineering. (p. 18). Nantes, France: ASME.
Miedema, S. A. (2013). Constructing the Shields Curve: Part C Cohesion by Silt, Hjulstrom, Sundborg. OMAE (p.
22). Nantes: ASME.
Miedema, S. A. (2013S). Software MS Excel 2LM & 3LM. Retrieved from The Delft Head Loss & Limit Deposit
Velocity Model: www.dhlldv.com
Miedema, S. A. (2014). An analytical approach to explain the Fuhrboter equation. Maritime Engineering, Vol.
167, Issue 2., 68-81.

Page 290 of 304 TOC Copyright © Dr.ir. S.A. Miedema


References.

Miedema, S. A. (2014). Dredging Processes Hydraulic Transport. Delft, Netherlands: Delft University of
Technology.
Miedema, S. A. (2014). The Delft Sand, Clay & Rock Cutting Model. (1st ed.). Delft: IOS Press, Delft University
Press. doi:10.3233/978-1-61499-454-1-i
Miedema, S. A. (2014B). An analysis of slurry transport at low line speeds. ASME 2014 33rd International
Conference on Ocean, Offshore and Arctic Engineering, OMAE. (p. 11). San Francisco, USA.: ASME.
Miedema, S. A. (2014C). An analytical approach to explain the Fuhrboter equation. Maritime Engineering, Vol.
167(2)., 1-14.
Miedema, S. A. (2014W). DHLLDV/Experiments. Retrieved from The Delft Head Loss & Limit Deposit Velocity
Model.: www.dhlldv.com
Miedema, S. A. (2015). A head loss model for homogeneous slurry transport. Journal of Hydrology &
Hydrodynamics, Vol. 63(1)., 1-12.
Miedema, S. A. (2015). Head Loss Model for Slurry Transport in the Heterogeneous Regime. Submitted to the
Journal of Ocean Engineering.
Miedema, S. A. (2015A). A head loss model for homogeneous slurry transport. Journal of Hydrology &
Hydrodynamics, Vol. 1., 14 pages.
Miedema, S. A. (2015B). Head Loss Model for Slurry Transport in the Heterogeneous Regime. Journal of Ocean
Engineering. Accepted.
Miedema, S. A. (2015B). THE SLIP RATIO OR HOLDUP FUNCTION IN SLURRY TRANSPORT. Dredging
Summit and Expo 2015. (p. 12). Houston, Texas, USA.: WEDA.
Miedema, S. A. (2015C). Head Loss Model for Slurry Transport in the Heterogeneous Regime. Journal of Ocean
Engineering.
Miedema, S. A., & Becker, S. (1993). The Use of Modeling and Simulation in the Dredging Industry, in Particular
the Closing Process of Clamshell Dredges. CEDA Dredging Days (p. 26 pages). Amsterdam, The
Netherlands: CEDA.
Miedema, S. A., & Frijters, D. (2003). The Mechanism of Kinematic Wedges at Large Cutting Angles - Velocity
and Friction Measurements. 23rd WEDA Technical Conference/35th TAMU Dredging Seminar (p. 14
pages). Chicago, Illinois, USA: WEDA/TAMU.
Miedema, S. A., & Frijters, D. (2004). The wedge mechanism for cutting of water saturated sand at large cutting
angles. WODCON XVII. Hamburg, Germany: WODA.
Miedema, S. A., & He, J. (2002B). The Existance of Kinematic Wedges at Large Cutting Angles. WEDA XXII
Technical Conference/34th Texas A/M Dredging Seminar (p. 20 pages). Denver, Colorado, USA:
WEDA/TAMU.
Miedema, S. A., & Ma, Y. (2002A). The Cutting of Water Saturated Sand at Large Cutting Angles. ASCE Dredging
02 (p. 16 pages). Orlando, Florida, USA: ASCE.
Miedema, S. A., & Matousek, V. (2014). An explicit formulation of bed friction factor for sheet flow. International
Freight Pipeline Society Symposium, 15th. (p. 17 pages). Prague, Czech Republic: IFPS.
Miedema, S. A., & Ramsdell, R. C. (2011). Hydraulic transport of sand/shell mixtures in relation with the critical
velocity. Terra et Aqua, Vol. 122.
Miedema, S. A., & Ramsdell, R. C. (2013). A head loss model for slurry transport based on energy considerations.
World Dredging Conference XX (p. 14). Brussels, Belgium: WODA.
Miedema, S. A., & Ramsdell, R. C. (2014). An Analysis of the Hydrostatic Approach of Wilson for the Friction
of a Sliding Bed. WEDA/TAMU (p. 21). Toronto, Canada: WEDA.
Miedema, S. A., & Ramsdell, R. C. (2014A). The Delft Head Loss & Limit Deposit Velocity Model.
Hydrotransport (p. 15). Denver, USA.: BHR Group.
Miedema, S. A., & Ramsdell, R. C. (2014B). An Analysis of the Hydrostatic Approach of Wilson for the Friction
of a Sliding Bed. WEDA/TAMU (p. 21). Toronto, Canada: WEDA.
Miedema, S. A., & Ramsdell, R. C. (2015). The Limit Deposit Velocity Model, a New Approach. Journal of
Hydrology & Hydromechanics, submitted., 15.
Miedema, S. A., & Vlasblom, W. J. (1996). Theory for Hopper Sedimentation. 29th Annual Texas A&M Dredging
Seminar & WEDA Conference. (p. 10). New Orleans, USA.: WEDA.
Miedema, S. A., & Yi, Z. (2001). An Analytical Method of Pore Pressure Calculations when Cutting Water
Saturated Sand. Texas A/M 33nd Annual Dredging Seminar (p. 18 pages). Houston, USA:
WEDA/TAMU.
Miedema, S. A., & Zijsling, D. (2012). Hyperbaric rock cutting. OMAE International Conference on Ocean,
Offshore and Arctic Engineering (p. 14). Rio de Janeiro, Brazil: ASME.
Miedema, S. A., Riet, E. J., & Matousek, V. (2002). Theoretical Description And Numerical Sensitivity Analysis
On Wilson Model For Hydraulic Transport Of Solids In Pipelines. WEDA Journal of Dredging
Engineering.

Copyright © Dr.ir. S.A. Miedema TOC Page 291 of 304


Introduction Dredging Engineering.

Miedema, S., & Rhee, C. v. (2007). A sensitivity analysis on the effects of dimensions and geometry of Trailing
Suction Hopper Dredges. WODCON. Orlando, Florida, USA: WODA.
Miedema, S., & Vlasblom, W. (1996). Theory of Hopper Sedimentation. 29th Annual Texas A&M Dredging
Seminar. New Orleans: WEDA/TAMU.
Miller, D., & Bruggers, D. (1980). Soil and permafrost conditions in the Alaskan Beaufort Sea. OTC 1980.
Houston, Texas, USA.: OTC.
Ming, G., Ruixiang, L., Ni, F., & Liqun, X. (2007). Hydraulic transport of coarse gravel. WODCON XVIII.
Orlando, Florida, USA: WODA.
Mitchell, J. (1976). Fundamentals of soil behavior. John Wiley & Sons, Inc.
Mitchell, J., Campanella, R., & Singh, A. (1968). Soil creep as a rate process. Journal SMFD, vol. 94, no. 1, ASCE.
Mogi, K. (1966). Pressure dependence of rock strength and transition from brittle fracture to ductile flow. Bulletin
Earthquake Res. Inst. Japan, Vol. 44., 215-232.
Moody, L. F. (1944). Friction Factors for Pipe Flow. Transactions of the ASME 66 (8)., 671-684.
Morsi, S., & Alexander, A. (1972). An investigation of particle trajectories in two-phase flow systems. Journal of
Fluid Mechanics, Vol. 55, 193-208.
Mukhtar, A. (1991). Investigations of the flow of multisized heterogeneous slurries in straight pipe and pipe bends.
Delhi, India: PhD Thesis, IIT.
Newitt, D. M., Richardson, M. C., Abbott, M., & Turtle, R. B. (1955). Hydraulic conveying of solids in horizontal
pipes. Transactions of the Institution of Chemical Engineers Vo.l 33., 93-110.
Nezu, I., & Nakagawa, H. (1993). Turbulence in Open Channel Flows. A. A. Balkema.
Ni, F., Zhao, L., Matousek, V., Vlasblom, W. J., & Zwartbol, A. (2004). Two phase flow of highly concentrated
slurry in a pipeline. Journal of Hydrodynamics, Series B, Vol. 16, No. 3., 325-331.
Ni, F., Zhao, L., Xu, L., & Vlasblom, W. J. (2008). A model calculation for flow resistance in the hydraulic
transport of sand. WODCON 18 (pp. 1377-1384). Orlando, Florida, USA: WODA.
Nielsen, P. (1981). Dynamics and geometry of wave generated ripples. Journal of Geophysics Research, Vol. 86.,
6467-6472.
Nikuradse, J. (1933, July/August). Stromungsgesetze in rauen Rohren. VDI Forschungsheft 361, Beilage zo
"Forschung auf dem Gebiete des Ingenieurwesens", Ausgabe B, Band 4.
Nishimatsu, Y. (1972). The mechanics of rock cutting. International Journal of Rock Mechanics & Mining
Science, vol. 9., 261-270.
Nnadi, F. N., & Wilson, K. C. (1992). Motion of contact load particles at high shear stress. Journal of Hydraulic
Engineering, Vol. 118., 1670-1684.
Nnadi, F. N., & Wilson, K. C. (1995). Bed Load Motion at High Shear Stress: Dune Washout and Plane Bed Flow.
Journal of Hydraulic Engineering, Vol. 121., 267-273.
O'Brien, M. P. (1933). Review of the theory of turbulent flow and its relations to sediment transportation.
Transactions of the American Geophysics Union, Vol. 14., 487-491.
O'Brien, M. P., & Folsom, R. G. (1939). The transportation of sand in pipelines. Vol. 3,No. 7 of University of
California publications in engineering.
Ooijens, S. (1999). Adding Dynamics to the Camp Model for the Calculation of Overflow Losses. Terra et Aqua
76, 12-21.
Ooijens, S., Gruijter, A. d., Nieuwenhuizen, A., & Vandycke, S. (2001). Research on Hopper Settlement Using
Large Scale Modeling. CEDA Dredging Days 2001 (pp. 1-11). Rotterdam: CEDA.
Oroskar, A. R., & Turian, R. M. (1980). The hold up in pipeline flow of slurries. AIChE, Vol. 26., 550-558.
Os, A. G. (1976). Snelle deformatie van korrelvormig materiaal onder water. PT-P31, no. 12, 735-741.
Os, A. G. (1977A). Behavior of soil when excavated under water. In International course on modern dredging.
The Hague, Netherlands.
Os, A. G. (1977B). Snelle deformatie van korrelvormig materiaal onder water. PT-B32, no. 8., 461-467.
Osman, M. (1964). The mechanics of soil cutting blades. J.A.E.R. 9 (4), 313-328.
Osterkamp, T. (2001). Sub-Sea Permafrost, Encyclopaedia of Ocean Sciences (2902-2912).
Palmer, A. (1999). Speed effects in cutting and ploughing. Geotechnique 49, no. 3., 285-294.
Paphitis, D. (2001). Sediment movement under unidirectional flows: an assesment of empirical threshold curves.
Coastal Engineering, 227-245.
Parzonka, W., Kenchington, J. M., & Charles, M. E. (1981). Hydrotransport of solids in horizontal pipes: Effects
of solids concentration and particle size on the deposit velocity. Canadian Journal of Chemical
Engineering, Vol. 59., 291-296.
Patankar, S. (1980). Numerical heat transfer and fluid flow. New York, USA: McGraw-Hill.
Peker, S. M., & Helvaci, S. S. (2008). Solid-Liquid Two Phase Flow. Amsterdam, The Netherlands: Elsevier.
Poloski, A. P., Etchells, A. W., Chun, J., Adkins, H. E., Casella, A. M., Minette, M. J., & Yokuda, S. (2010). A
pipeline transport correlation for slurries with small but dense particles. Canadian Journal of Chemical
Engineering, Vol. 88., 182-189.

Page 292 of 304 TOC Copyright © Dr.ir. S.A. Miedema


References.

Postma, H. (1967). Sediment transport and sedimentation in the estuarine environment. Estuaries, AAAS,
Washington D.C. Publ. 83., 158-179.
Prandl, L. (1925). Z. angew. Math. Mech. 5 (1), 136-139.
Pugh, F. J., & Wilson, K. C. (1999). Role of the interface in stratified slurry flow. Powder Technology, Vol. 104.,
221-226.
Pugh, F. J., & Wilson, K. C. (1999). Velocity and concentration distribution in sheet flow above plane beds.
Journal of Hydraulic Engineering, 117-125.
Raalte, G., & Zwartbol, A. (1986). Disc bottom cutterhead, a report on laboratory and field tests. WODCON XI.
Brighton, UK: WODA.
Rafatian, N., Miska, S., Ledgerwood III, L., Hughes, B., Ahmed, R., Yu, M., & Takach, N. (2009). Experimental
study of MSE of a single PDC cutter under simulated pressurized conditions. SPE/IADC 119302 Drilling
Conference & Exhibition. Amsterdam, Netherlands: SPE International.
Ramsdell, R. C., & Miedema, S. A. (2010). Hydraulic transport of sand/shell mixtures. WODCON XIX. Beijing,
China.: WODA.
Ramsdell, R. C., & Miedema, S. A. (2013). An overview of flow regimes describing slurry transport. WODCON
XX (p. 15). Brussels, Belgium.: WODA.
Ramsdell, R. C., Miedema, S. A., & Talmon, A. (2011). Hydraulic transport of sand/shell mixtures. OMAE 2011.
Rotterdam, Netherlands.: ASME.
Randal, R. E., Jong, P. S., & Miedema, S. A. (2000). Experiences with Cutter Suction Dredge Simulator Training.
WEDA/TAMU (p. 10). Rhode Island, USA: WEDA.
Raudviki, A. J. (1990). Loose Boundary Hydraulics. University of Auckland: Pergamon Press.
Ravelet, F., Bakir, F., Khelladi, S., & Rey, R. (2012). Experimental study of hydraulic transport of large particles
in horizontal pipes. Experimental Thermal and Fluid Science, 13.
Reece, A. (1965). The fundamental equation of earth moving machinery. Symposium Earth Moving Machinery.
London: Institute of Mechanical Engineering.
Reichardt, H. (1951). Vollstandige Darstellung der Turbulenten Geswindigkeitsverteilung in Glatten Leitungen.
Zum Angew. Math. Mech., 3(7), 208-219.
Rhee, C. v. (2002A). The influence of the bottom shear stress on the sedimentation of sand. 11th International
Symposium on Transport and Sedimentation of Solid Particles. Ghent, Belgium.
Rhee, C. v. (2002B). Numerical modeling of the flow and settling in a Trailing Suction Hopper Dredge. 15th
International Conference on Hydrotransport. Banff, Canada.
Rhee, C. v. (2002C). On the sedimentation process in a Trailing Suction Hopper Dredger. Delft, Netherlands:
Delft University of Technology, PhD Thesis.
Rhee, C., & Steeghs, H. (1991 June). Multi blade ploughs in saturated sand, model cutting tests. Dredging & Port
Construction.
Richardson, J. F., & Zaki, W. N. (1954). Sedimentation & Fluidization: Part I. Transactions of the Institution of
Chemical Engineering 32, 35-53.
Riet, E. J., Matousek, V., & Miedema, S. A. (1995). A Reconstruction of and Sensitivity Analysis on the Wilson
Model for Hydraulic Particle Transport. Proc. 8th Int. Conf. on Transport and Sedimentation of Solid
Particles. Prague, Czech Republic.
Riet, E. J., Matousek, V., & Miedema, S. A. (1996). A Theoretical Description and Numerical Sensitivity Analysis
on Wilson's Model for Hydraulic Transport in Pipelines. Journal of Hydrology & Hydromechanics.
Rijn, L. v. (1984). Sediment transport: Part I: Bed load transport. Journal of Hydraulic Engineering, Vol. 110(10),
1431-1456.
Rijn, L. v. (1993). Principles of Sediment Transport, Part 1. . Blokzijl: Aqua Publications.
Roberts, J., Jepsen, R., Gotthard, D., & Lick, W. (1998). Effects of particle size and bulk density on erosion of
quartz particles. Journal of Hydraulic Engineering, 1261-1267.
Robinson, M. P. (1971). Critical deposit velocities for low concentration solid-liquid mixtures. MSc Thesis. Lehigh
University, Fritz Laboratory.
Robinson, M. P., & Graf, W. H. (1972). Critical deposit velocities for low concentration sand water mixtures.
ASCE National Water Resources EnVg Meeting Preprint 1637, January 1972. Paper 1982. Atlanta,
Georgia, USA.: Lehigh University, Fritz Laboratory.
Rodi, W. (1993). Turbulence models and their application in hydraulics, a state of the art review. IAHR, Third
Edition.
Rouse, H. (1937). Modern conceptions of the mechanics of fluid turbulence. Transactions of the American Society
of Cicil Engineers, Vol. 102, 463-505, Discussion 506-543.
Rowe, P. (1962). The stress dilatancy relation for static equilibrium of an assembly of particles in contact.
Proceedings Royal Society A269. (pp. 500-527). Royal Society.
Rowe, P. (1987). A convinient empirical equation for estimation of the Richardson-Zaki exponent. Chemical
Engineering Science Vol. 42, no. 11, 2795-2796.

Copyright © Dr.ir. S.A. Miedema TOC Page 293 of 304


Introduction Dredging Engineering.

Rowe, P. N. (1987). A convinient empirical equation for estimation of the Richardson-Zaki exponent. Chemical
Engineering Science Vol. 42, no. 11, 2795-2796.
Roxborough, F. (1987). The role of some basic rock properties in assessing cuttability. Seminar on Tunnels -
Wholly Engineered Structures (pp. 1-21). Canberra, Australia: AFCC.
Saffman, P. G. (1965). The lift on small sphere in a slow shear low. Journal of Fluid Mechanics, 22, 385-400.
Sanders, R. S., Sun, R., Gillies, R. G., McKibben, M., Litzenberger, C., & Shook, C. A. (2004). Deposition
Velocities for Particles of Intermediate Size in Turbulent Flows. Hydrotransport 16 (pp. 429-442).
Santiago, Chile.: BHR Group.
Schaan, J., & Shook, C. A. (2000). Anomalous friction in slurry flows. Canadian Journal of Chemical
Engineering, Vol. 78., 726-730.
Scheurel, H. G. (1985). Rohrverschleiss beim hydraulischen feststoffentransport. Karlsruhe: Universitat Karlsruhe.
Schiller, R. E., & Herbich, J. B. (1991). Sediment transport in pipes. Handbook of dredging. New York: McGraw-
Hill.
Schlichting, H. (1968). Boundary layer theory. 6th ed. New York: McGraw-Hill.
Schrieck, G. (1996). Introduction to Dredging Engineering. Delft, the Netherlands: Delft University of
Technology.
Segal, G. (2001). Sepra analysis programmers guide, standard problems and users manual. Leidschendam,
Netherlands: Ingenieursbureau Sepra.
Sellgren, A., & Wilson, K. (2007). Validation of a four-component pipeline friction-loss model. Hydrotransport
17 (pp. 193-204). BHR Group.
Seshadri, V., Singh, S. N., & Kaushal, D. R. (2006). A model for the prediction of concentration and particle size
distribution for the flow of multisized particulate suspensions through closed ducts and open channels.
Particulate Science and Technology: An International Journal., 239-258.
Shields, A. (1936). Anwendung der Aehnlichkeitsmechanik und der Turbulenzforschung auf die
Geschiebebewegung. Mitteilung der Preussischen Versuchsanstalt fur Wasserbau und Schiffbau, Heft 26,
Berlin. Belin.
Shook, C. A., Geller, L., Gillies, R. G., Husband, W. H., & Small, M. (1986). Experiments with coarse particles
in a 250 mm pipeline. 10th International Conference on the Hydraulic Transport of Solids in Pipelines
(Hydrotransport 10). (pp. 219-227). Cranfield, UK.: BHRA Fluid Eng.
Shook, C. A., Gillies, R. G., & Sanders, R. S. (2002). Pipeline Hydrotransport with Application in the Oil Sand
Industry. Saskatoon, Canada: Saskatchewan Research Council, SRC Publication 11508-1E02.
Shook, C., & Roco, M. (1991). Slurry Flow, Principles & Practice. Boston: Butterworth Heineman.
Silin, M. O., Kobernik, S. G., & Asaulenko, I. A. (1958). Druckhohenverluste von Wasser und Wasser-Boden-
Gemischen in Rohrleitungen grossen Durchmessers. Ukrain: Dopovidi Akad. Nauk. Ukrain RSR.
Silin, N. A., & Kobernik, S. G. (1962). Rezimy raboty zemlijesosnych snarjadov, Kijev.
Simons, D. (1957). Theory and design of stable channels in alluvial material. PhD thesis: Colorado State
University.
Sinclair, C. G. (1962). The limit deposit velocity of heterogeneous suspensions. Proceedings Symposium on the
Interaction Btween Fluids and Particles. Institute of Chemical Engineers.
Smoldyrev. (1970). Truboprowodnyi transport( rohrleitungstransport). Moskau.
Sobota, J., & Kril, S. I. (1992). Liquid and solid velocity during mixture flow. Proceedings 10th International
Colloquium Massenguttransport durch Rohrleitungen., (p. K). Meschede, Germany.
Sohne, W. (1956). Some basic considerations of soil mechanics as applied to agricultural engineering. Grundlagen
der landtechnik (7)., 11-27.
Soleil, G., & Ballade, P. (1952). Le transport hydraulique des materiaux dans les travaux publics, observations des
resultats d'essais en grandeur nature. Deuxiemes Journees de l'Hydraulique, 9-26.
Soulsby, R., & Whitehouse, R. (1997). Threshold of sediment motion in coastal environment. Proceedings Pacific
Coasts and Ports. (pp. 149-154). Christchurch, New Zealand: University of Canterbury.
Souza Pinto, T. C., Moraes Junior, D., Slatter, P. T., & Leal Filho, L. S. (2014). Modelling the critical velocity for
heterogeneous flow of mineral slurries. International Journal of Multiphase Flow, 65., 31-37.
Soydemir, C. (1977). Potential models for shear strength generation in soft marine clays. International Symposium
on Soft Clay. Bangkok, Thailand.
Spelay, R., Hashemi, S. A., Gillies, R. G., Hegde, R., Sanders, R. S., & Gillies, D. G. (2013). Governing friction
loss mechanisms and the importance of offline characterization tests in the pipeline transport of dense
coarse particle slurries. Proceedings of the ASME 2013 Fluids Engineering Division Summer Meeting.
(pp. 1-7). Incline Village, Nevada, USA.: FEDSM2013.
Stam, P. (1983). Analyse ten behoeve van het ontwerp van een klei snijdende sleepkop, CO/82/129. Delft,
Netherlands: Delft University of Technology.
Stansby, P. (1997). Semi-implicit finite shallow-water flow and solute transport solver with k-epsilon turbulence
model. International Journal for Numerical Methods in Fluids, vol. 25., 285-313.

Page 294 of 304 TOC Copyright © Dr.ir. S.A. Miedema


References.

Steeghs, H. (1985A). Snijden van zand onder water, part I. Ports & Dredging no. 121.
Steeghs, H. (1985B). Snijden van zand onder water, part II. Ports & Dredging no. 123.
Stelling, G., & Kester, J. (1994). On the approximation of horizontal gradients in sigma coordinates for bathymetry
with steep bottom slopes. International Journal for Numerical Methods in Fluids, vol. 18, 915-935.
Stevenson, P., Cabrejos, F. J., & Thorpe, R. B. (2002). Incipient motion of particles on a bed of like particles in
hydraulic and pneumatic conveying. Fourth World Congress of Particle Technology, Sydney, 21st–25th
July (paper 400). Sydney.
Stevenson, P., Thorpe, R. B., & Davidson, J. F. (2002). Incipient motion of a small particle in the viscous boundary-
layer at a pipe wall. Chemical Engineering Science, 57, 4505–4520.
Sundborg, A. (1956). The River Klarålven: Chapter 2. The morphological activity of flowing water erosion of the
stream bed. Geografiska Annaler, 38, 165-221.
Swamee, P. K. (1993). Critical depth equations for irrigation canals. Journal of Irrigation and Drainage
Engineering, ASCE., 400-409.
Swamee, S. E., & Jain, K. A. (1976). Explicit equations for pipe-flow problems. Journal of the Hydraulics Division
(ASCE) 102 (5)., 657-664.
Talmon, A. (2013). Analytical model for pipe wall friction of pseudo homogeneous sand slurries. Particulate
Science & technology: An International Journal, 264-270.
Talmon, A. M. (2011). Hydraulic Resistance of Sand-Water Mixture Flow in Vertical Pipes. T&S, Transport and
Sedimentation of Solid Particles (pp. 137-147). Wroclaw, Poland: T&S.
Talmon, A. M. (2013). Analytical model for pipe wall friction of pseudo homogeneous sand slurries. Particulate
Science & technology: An International Journal, 264-270.
Televantos, Y., Shook, C. A., Carleton, A., & Street, M. (1979). Flow of slurries of coarse particles at high solids
concentrations. Canadian Journal of Chemical Engineering, Vol. 57., 255-262.
Terzaghi, K., & Peck, R. (1964). Soil mechanics in engineering practise. New York: John Wiley & Sons.
Thomas, A. (1976). SCALE-UP METHODS FOR PIPELINE TRANSPORT OF SLURRIES. International
Journal of Mineral Processing, Vol. 3., 51-69.
Thomas, A. D. (1979). Predicting the deposit velocity for horizontal turbulent pipe flow of slurries. International
Journal of Multiphase Flow, Vol. 5., 113-129.
Thomas, A. D. (2014). Slurries of most interest to the mining industry flow homogeneously and the deposit
velocity is the key parameter. HydroTransport 19. (pp. 239-252). Denver, Colorado, USA.: BHR Group.
Thomas, D. G. (1962). Transport Characteristics of Suspensions: Part VI. Minimum velocity for large particle size
suspensions in round horizontal pipes. A.I.Ch.E. Journal, Vol.8(3)., 373-378.
Thomas, D. G. (1965). Transport characteristics of suspensions: VIII. A note on the viscosity of Newtonian
suspensions of uniform spherical particles. Journal Of Colloidal Sciences, Vol. 20., 267-277.
Turian, R. M., & Yuan, T. F. (1977). Flow of slurries in pipelines. AIChE Journal, Vol. 23., 232-243.
Turian, R. M., Hsu, F. L., & Ma, T. W. (1987). Estimation of the critical velocity in pipeline flow of slurries.
Powder Technology, Vol. 51., 35-47.
Turnage, G., & Freitag, D. (1970). Effects of cone velocity and size on soil penetration resistance. ASEA 69-670.
Turner, T. (1996). Fundamentals of Hydraulic Dredging. New York: ASCE.
Turton, R., & Levenspiel, O. (1986). A short note on the drag correlation for spheres. Powder technology Vol. 47,
83-85.
Vanoni, V. A. (1975). Sedimentation Engineering: American Society of Civil Engineers, Manuals and Reports on
Engineering Practice. No. 54. P.745.
Verhoef, P. (1997). Wear of rock cutting tools: Implications for site investigation of rock dredging projects. Delft,
Netherlands: Balkema Rotterdam.
Verruijt, A. (1983). Soil Mechanics. Delft: DUM, Netherlands.
Vlasak, P. (2008). Laminar, transitional and turbulent flow of fine grained slurries in pipelines. Prague.: Czech
Technical University in Prague, Fakulty of Civil Engineering.
Vlasak, P., Kysela, B., & Chara, Z. (2012). FLOW STRUCTURE OF COARSE-GRAINED SLURRY IN A
HORIZONTAL PIPE. Journal of Hydrology & Hydromechanics, Vol. 60., 115-124.
Vlasblom, W. (2003-2007). Rock Cutting, Lecture Notes. Delft, Netherlands: Delft University of Technology.
Vlasblom, W., & Miedema, S. (1995). A Theory for Determining Sedimentation and Overflow Losses in Hoppers.
WODCON IV. Amsterdam, Netherlands: WODA.
Vocadlo, J. J. (1972). Prediction of pressure gradient for the horizontal turbulent flow of slurries. Hydrotransport
2. Coventry: BHRA.
Vocadlo, J. J., & E., C. M. (1972). Prediction of pressure gradient for the horizontal turbulent flow of slurries.
Conference on the Hydraulic Transport of Solids in Pipes. Warwick, England: British Hydromechanics
Research Association.
Vukovic, M., & Soro, A. (1992). Determination of hydraulic conductivity of porous media from grain size
composition. Water Resources Publications, Littleton, Colorado.

Copyright © Dr.ir. S.A. Miedema TOC Page 295 of 304


Introduction Dredging Engineering.

Wallis, G. (1969). One Dimensional Two Phase Flow. McGraw Hill.


Wasp, E. J. (1963). Cross country coal pipeline hydraulics. Pipeline News., 20-28.
Wasp, E. J., & Slatter, P. T. (2004). Deposition velocities for small particles in large pipes. 12th International
Conference on Transport & Sedimentation of Solid Particles, (pp. 20-24). Prague, Czech Republic.
Wasp, E. J., Kenny, J. P., & Gandhi, R. L. (1977). Solid liquid flow slurry pipeline transportation. Transactions
Technical Publications.
Wasp, E. J., Kenny, J. P., Aude, T. C., Seiter, R. H., & Jacques, R. B. (1970). Deposition velocities transition
velocities and spatial distribution of solids in slurry pipelines. Hydro Transport 1, paper H42. (pp. 53-
76). Coventry: BHRA Fluid Engineering.
Weibull, W. (1939). A statistical theory of the strength of materials. Royal Swedish Institute of Engineers, 151:1.
Weijermars, R. (1997-2011). Principles of rock mechanics. Delft, Netherlands: Alboran Science Publishing.
Welte, A. (1971). Grundlagen der Berechnung der Rohrleitungsdruckverluste. Konstruction 23, Heft 5 & 6.
Westendorp, J. H. (1948). Verslag literatuurstudie over persen van zand M.276. Delft, Netherlands: Delft
Hydraulics Laboratory.
White, C. M. (1940). The equilibrium of grains on the bed of a stream. Proceedings Royal Society of London,
A174, pp. 322-338.
Whitlock, L., Wilson, K. C., & Sellgren, A. (2004). Effect of near-wall lift on frictional characteristics of sand
slurries. Hydrotransport 16 (pp. 443-454). Cranfield, UK.: BHR Group.
Wiberg, P. L., & Smith, J. D. (1987A). Calculations of the critical shear stress for motion of uniform and
heterogeneous sediments. Water Resources Research, 23(8), 1471–1480.
Wiberg, P., & Smith, J. (1987B). Initial motion of coarse sediment in streams of high gradient. Proceedings of the
Corvallis Symposium. IAHS Publication No. 165.
Wiedenroth, W. (1967). Untersuchungen uber die forderung von sand wasser gemischen durch rohrleitungen und
kreiselpumpen. Hannover: PhD Thesis, Technische Hochschule Hannover.
Wikramanayake, P. N., & Madsen, O. S. (1991). Calculation of movable bed friction factors. Vicksburg,
Mississippi.: Tech. Rep. DACW-39-88-K-0047, 105 pp., Coastal Eng. Res. Cent.,.
Wilson, K. C. (1965). Application of the minimum entropy production principle to problems in two-phase flow,
PhD Thesis. Kingston, Ontario, Canada.: Queens University.
Wilson, K. C. (1966). Bed Load Transport at High Shear Stress. Journal of the Hydraulics Division, 49-59.
Wilson, K. C. (1970). Slip point of beds in solid liquid pipeline flow. Journal of Hydraulic Division, Vol 96(HY1),
1-12.
Wilson, K. C. (1970). Slip point of beds in solid-liquid pipeline flow. Proceedings American Society of Civil
Engineers, Vol. 96, HY1.
Wilson, K. C. (1972). A Formula for the Velocity Required to Initiate Particle Suspension in Pipeline Flow.
Hydrotransport 2 (pp. E2 23-36). Warwick, UK.: BHRA Fluid Engineering.
Wilson, K. C. (1974). Coordinates for the Limit of Deposition in Pipeline Flow. Hydrotransport 3 (pp. E1 1-13).
Colorado School of Mines, Colorado, USA.: BHRA Fluid Engineering.
Wilson, K. C. (1975). Stationary Deposits and Sliding Beds in Pipes Transporting Solids. Dredging Technology
(pp. C3 29-40). College Station, Texas, USA.: BHRA Fluid Engineering.
Wilson, K. C. (1976). A Unified Physically based Analysis of Solid-Liquid Pipeline Flow. Hydrotransport 4 (pp.
A1 1-16). Banff, Alberta, Canada: BHRA Fluid Engineering.
Wilson, K. C. (1979). Deposition limit nomograms for particles of various densities in pipeline flow.
Hydrotransport 6 (p. 12). Canterbury, UK: BHRA Fluid Engineering.
Wilson, K. C. (1980). Analysis of Slurry Flows with a Free Surface. Hydrotransport 7 (pp. 123-132). Sendai,
Japan: BHRA Fluid Engineering.
Wilson, K. C. (1984). Analysis of Contact Load Distribution and Application to Deposition Limit in Horizontal
Pipes. Journal of Pipelines, Vol. 4., 171-176.
Wilson, K. C. (1986). Effect of Solids Concentration on Deposit Velocity. Journal of Pipelines, Vol. 5., 251-257.
Wilson, K. C. (1987). Analysis of Bed Load Motion at High Shear Stress. Journal of Hydraulic Engineering, Vol.
113., 97-103.
Wilson, K. C. (1988). Evaluation of interfacial friction for pipeline transport models. Hydrotransport 11 (p. B4).
BHRA Fluid Engineering.
Wilson, K. C. (1989). Mobile Bed Friction at High Shear Stress. Journal of Hydraulic Engineering, Vol. 115.,
825-830.
Wilson, K. C., & Addie, G. R. (1997). Coarse particle pipeline transport: effect of particle degradation on friction.
Powder Technology, Vol. 94., 235-238.
Wilson, K. C., & Brown, N. P. (1982). Analysis of Fluid Friction in dense Phase Pipeline Flow. The Canadian
Journal of Chemical Engineering, Vol. 60., 83-86.

Page 296 of 304 TOC Copyright © Dr.ir. S.A. Miedema


References.

Wilson, K. C., & Judge, D. G. (1976). New Techniques for the Scale-Up of Pilot Plant Results to Coal Slurry
Pipelines. Proceedings International Symposium on Freight Pipelines. (pp. 1-29). Washington DC, USA:
University of Pensylvania.
Wilson, K. C., & Judge, D. G. (1977). Application of Analytical Model to Stationary Deposit Limit in Sand Water
Slurries. Dredging Technology (pp. J1 1-12). College Station, Texas, USA.: BHRA Fluid Engineering.
Wilson, K. C., & Judge, D. G. (1978). Analytically based Nomographic Charts for Sand-Water Flow.
Hydrotransport 5 (pp. A1 1-12). Hannover, Germany: BHRA Fluid Engineering.
Wilson, K. C., & Judge, D. G. (1980). New Techniques for the Scale-up of Pilot Plant Results to Coal Slurry
Pipelines. Journal of Powder & Bulk Solids Technology., 15-22.
Wilson, K. C., & Nnadi, F. N. (1990). Behavior of Mobile Beds at High Shear Stress. Proceedings Coastal
Engineering 22., (pp. 2536-2541). Delft.
Wilson, K. C., & Pugh, F. J. (1988). Dispersive Force Basis for Concentration Profiles. Journal of Hydraulic
Engineering, Vol. 114, No. 7., 806-810.
Wilson, K. C., & Pugh, F. J. (1988). Dispersive Force Modelling of Turbulent Suspension in Heterogeneous Slurry
Flow. The Canadian Journal of Chemical Engineering, Vol. 66., 721-727.
Wilson, K. C., & Sellgren, A. (2001). Hydraulic transport of solids, Pump Handbook, pp. 9.321-9.349. McGraw-
Hill.
Wilson, K. C., & Sellgren, A. (2003). Interaction of Particles and Near-Wall Lift in Slurry Pipelines. Journal of
Hydraulic Engineering, Vol. 129., 73-76.
Wilson, K. C., & Sellgren, A. (2010). Behavior of intermediate particle slurries in pipelines. Hydrotransport 18
(pp. 117-128). Rio de Janeiro: BHR Group.
Wilson, K. C., & Sellgren, A. (2012). Revised Method for Calculating Stratification Ratios for Heterogeneous
Flows. 14th International Conference on Transport & Sedimentation of Solid Particles., (pp. 334-340).
Wilson, K. C., & Watt, W. E. (1974). Influence of Particle Diameter on the Turbulent Support of Solids in Pipeline
Flow. Hydrotransport 3 (pp. D1 1-9). Colorado School of Mines, Colorado, USA.: BHRA Fluid
Engineering.
Wilson, K. C., Addie, G. R., & Clift, R. (1992). Slurry Transport using Centrifugal Pumps. New York: Elsevier
Applied Sciences.
Wilson, K. C., Addie, G. R., Clift, R., & Sellgren, A. (1997). Slurry Transport using Centrifugal Pumps. Glasgow,
UK.: Chapman & Hall, Blackie Academic & Professional.
Wilson, K. C., Addie, G. R., Sellgren, A., & Clift, R. (2006). Slurry transport using centrifugal pumps. New York:
Springer Science+Business Media Inc.
Wilson, K. C., Clift, R., & Sellgren, A. (2002). Operating points for pipelines carrying concentrated heterogeneous
slurries. Powder Technology, Vol. 123., 19-24.
Wilson, K. C., Clift, R., Addie, G. R., & Maffet, J. (1990). Effect of Broad Particle Grading on Slurry Stratification
Ratio and Scale-up. Powder Technology, 61., 165 - 172.
Wilson, K. C., Sanders, R. S., Gillies, R. G., & Shook, C. A. (2010). Verification of the near wall model for slurry
flow. Powder Technology, Vol. 197., 247-253.
Wilson, K. C., Sellgren, A., & Addie, G. R. (2000). Near-wall fluid lift of particles in slurry pipelines. 10th
Conference on Transport and Sedimentation of Solid Particles. Wroclav, Poland: T&S10.
Wilson, K. C., Streat, M., & Bantin, R. A. (1972). Slip model correlation of dense two phase flow. Hydrotransport
2 (pp. B1 1-10). Warwick, UK: BHRA Fluid Engineering.
Wilson, W. E. (1942). Mechanics of flow with non colloidal inert solids. Transactions ASCE Vol. 107., 1576-
1594.
Wismer, R., & Luth, H. (1972A). Performance of Plane Soil Cutting Blades. Transactions of ASEA.
Wismer, R., & Luth, H. (1972B). Rate effects in soil cutting. Journal of Terramechanics, vol. 8, no. 3., 11-21.
Wood, D. J. (1966). An explicit friction factor relationship. Civil Engineering, Vol. 36, ASCE., 60-61.
Worster, R. C., & Denny, D. F. (1955). Hydraulic transport of solid materials in pipelines. Institution of Mechanical
Engineers (London), 563-586.
Wu, W., & Wang, S. (2006). Formulas for sediment porosity and settling velocity. Journal of Hydraulic
Engineering, 132(8), 858-862.
Yagi, T. (1970). Sedimentation effects of soil in hopper. WODCON III (pp. 1-22). Singapore: WODA.
Yagi, T., Okude, T., Miyazaki, S., & Koreishi, A. (1972). An Analysis of the Hydraulic Transport of Solids in
Horizontal Pipes. Nagase, Yokosuka, Japan.: Report of the Port & Harbour Research Institute, Vol. 11,
No. 3.
Yalin, M. S., & Karahan, E. (1979). Inception of sediment transport. ASCE Journal of the Hydraulic Division,
105, 1433–1443.
Yi, Z. (2000). The FEM calculation of pore water pressure in sand cutting process by SEPRAN. Delft, Netherlands:
Delft University of Technology, Report: 2001.BT.5455.

Copyright © Dr.ir. S.A. Miedema TOC Page 297 of 304


Introduction Dredging Engineering.

Yi, Z., & Miedema, S. (2001). Finite Element Calculations To Determine The Pore Pressures When Cutting Water
Saturated Sand At Large Cutting Angles. CEDA Dredging Days (p. 20 pages). Amsterdam, The
Netherlands: CEDA.
Yi, Z., & Miedema, S. (2002). Finite Element Calculations To Determine The Pore Pressures When Cutting Water
Saturated Sand At Large Cutting Angles. CEDA Dredging Days (p. 20 pages). Amsterdam, The
Netherlands: CEDA.
Zandi, I. (1971). Hydraulic transport of bulky materials. In I. Zandi, Advances in Solid–Liquid Flow in Pipes and
its Applications. (pp. 1-38). Oxford: Pergamon Pres.
Zandi, I., & Govatos, G. (1967). Heterogeneous flow of solids in pipelines. Proc. ACSE, J. Hydraul. Div.,,
93(HY3)., 145-159.
Zanke, U. C. (1977). Berechnung der Sinkgeschwindigkeiten von Sedimenten. Hannover, Germany: Mitteilungen
Des Francius Instituts for Wasserbau, Heft 46, seite 243, Technical University Hannover.
Zanke, U. C. (2001). Zum Einfluss der Turbulenz auf den Beginn der Sedimentbewegung. Darmstadt, Germany:
Mitteilungen des Instituts fur Wasserbau und Wasserwirtschaft der TU Darmstadt, Heft 120.
Zanke, U. C. (2003). On the influence of turbulence on the initiation of sediment motion. International Journal of
Sediment Research, 18(1), 17–31.
Zeng, D., & Yao, Y. (1988). Investigation on the relationship between soil metal friction and sliding speed. 2nd
Asian Pacific Conference of ISTVS. Bangkok, Thailand.
Zeng, D., & Yao, Y. (1991). Investigation on the relationship between soil shear strength and shear rate. Journal
of Terramechanics, 28 (1).
Zijsling, D. (1987). Single cutter testing - a key for PDC bit development (SPE 16529). Offshore Europe 87.
Aberdeen, Scotland.

Page 298 of 304 TOC Copyright © Dr.ir. S.A. Miedema


List of Figures.

Chapter 10: List of Figures.


Figure 1-1: Earthwork in Germany (source Wikimedia). ....................................................................................... 2
Figure 1-2: Fox glacier, New Zealand (source Wikimedia). ................................................................................... 3
Figure 1-3: Soil naming according to USDA. ......................................................................................................... 4
Figure 1-4: Soil failure (www.4isfge.org). .............................................................................................................. 5
Figure 1-5: The Wenjiagou landslide (blogs.agu.org)............................................................................................. 6
Figure 1-6: Karl von Terzaghi, one of the founders of modern soil mechanics. ..................................................... 6
Figure 1-7: Sand from the Gobi desert, Mongolia (source Wikimedia). ................................................................. 7
Figure 1-8: Sand in the Sahara desert (source Luca Galuzzi – www.galuzzi.it) ..................................................... 8
Figure 1-9: Quaternary clay in Estonia (source Wikimedia) ................................................................................... 9
Figure 1-10: Varved clay, Little River State Park, Waterbury, Vermont (source www.anr.state.vt.us). ..............10
Figure 1-11: Sample of igneous gabbro, Rock Creek Canyon, California (source Wikimedia). ...........................11
Figure 1-12: Sandstone formations, Vermillion Cliffs, Arizona (source www.reddit.com). .................................13
Figure 1-13: Columns of Basalt of the Scottish Island of Staffa (National Geographic). ......................................13
Figure 1-14 A: Aid to identification of rock for engineering purposes (After BS 5930:1981). ............................14
Figure 1-15 B: Aid to identification of rock for engineering purposes (After BS 5930:1981). ............................15
Figure 1-16: Utica Shale, Fort Plain, New York (Wikipedia). ...............................................................................16
Figure 1-17: The rock formation cycle (galleryhip.com). ......................................................................................16
Figure 1-18: The particle size distributions of the sands used by Roberts et al. (1998). .......................................17
Figure 1-19: Liquid limit device. ...........................................................................................................................18
Figure 1-20: Liquid limit device. ...........................................................................................................................18
Figure 1-21: The relation between SL, PL, LL and PI. ..........................................................................................19
Figure 1-22: SPT values versus relative density (Miedema (1995). ......................................................................22
Figure 1-23: Friction angle versus SPT value (Miedema (1995). ..........................................................................27
Figure 1-24: A UCS test facility (Timely Engineering Soil Tests, LLC). ..............................................................30
Figure 1-25: Bending (Vlasblom (2003-2007)). ...................................................................................................30
Figure 1-26: The Brazilian split test (Vlasblom (2003-2007)). ..............................................................................31
Figure 1-27: Diagram showing definitions and directions for Darcy’s law. ..........................................................36
Figure 1-28: Brittle failure types (Vlasblom (2003-2007)). ...................................................................................38
Figure 1-29: Brittle-ductile failure of marble (M.S. Patterson, Australian National University). ..........................38
Figure 1-30: A set of sieves (Essa Australia from: www.directindustry.com). ......................................................39
Figure 1-31: The Standard Penetration Test (www.shalviengineering.com). ........................................................40
Figure 1-32: A typical CPT test setup (www.geotechdata.com). ...........................................................................42
Figure 1-33: Several configurations of cones (www.geotechdata.info). ................................................................42
Figure 1-34: Several cone configurations. .............................................................................................................43
Figure 1-35: The Triaxial apparatus (www.geotechdata.info). ..............................................................................44
Figure 1-36: The Triaxial apparatus cross-section (civilblog.org). ........................................................................45
Figure 1-37: The direct shear test. ..........................................................................................................................46
Figure 1-38: The vane shear test (English.geocpt.es).............................................................................................47
Figure 1-39: Shear vane and Torvane for soil testing (www.humboldtmfg.com). .................................................47
Figure 1-40: Point load test facility (inside.mines.edu). ........................................................................................48
Figure 1-41: Brazilian splitting tension test. ..........................................................................................................48
Figure 1-42: BTS zoomed. .....................................................................................................................................48
Figure 1-43: A BTS test after failure. ....................................................................................................................49
Figure 2-1: Forces on a settling particle. ................................................................................................................57
Figure 2-2: Experimental data for drag coefficients of spheres as a function of the Reynolds number (Turton &
Levenspiel, 1986). ...........................................................................................................................58
Figure 2-3: The particle Reynolds number as a function of the particle diameter. ................................................59
Figure 2-4: Drag coefficient as a function of the particle shape (Wu & Wang, 2006). ..........................................60
Figure 2-5: Drag coefficient for natural sediments (Sf=0.7) (Wu & Wang, 2006).................................................61
Figure 2-6: The drag coefficient as a function of the particle Reynolds number. ..................................................62
Figure 2-7: The settling velocity of individual particles. .......................................................................................63
Figure 2-8: The settling velocity of individual particles using the shape factor.....................................................65
Figure 2-9: The shape factor ξ as a function of the dimensionless particle diameter D*. ......................................67
Figure 2-10: The hindered settling power according to several researchers. .........................................................68
Figure 3-1: A pump –pipeline system. ...................................................................................................................77
Figure 3-2: The speed-power curve of a diesel engine. ..........................................................................................78
Figure 3-3: Centrifugal pumps. ..............................................................................................................................79
Figure 3-4: The pressure-flow curves. ...................................................................................................................80

Copyright © Dr.ir. S.A. Miedema TOC Page 299 of 304


Introduction Dredging Engineering.

Figure 3-5: The characteristics of the ladder pump. ...............................................................................................82


Figure 3-6: The characteristics of the main pump and the booster pump, torque limited. .....................................82
Figure 3-7: Characteristics of the pump/pipeline system, not limited. ...................................................................84
Figure 3-8: Characteristics of the pump/pipeline system, torque limited. ..............................................................84
Figure 3-9: The mass equilibrium in a pipe segment. ............................................................................................86
Figure 3-10: The system curves for 3 cases, accelerating. .....................................................................................87
Figure 3-11: The system curves for 3 cases, decelerating. .....................................................................................88
Figure 3-12: The density distribution in the pipeline after 12 minutes. .................................................................90
Figure 3-13: The density distribution in the pipeline after 17 minutes. .................................................................90
Figure 3-14: The density distribution in the pipeline after 22 minutes. .................................................................91
Figure 3-15: Line speed, density, total power and situ production as a function of time. ......................................92
Figure 3-16: Speed, power, vacuum and discharge pressure of the ladder pump vs. time. ....................................92
Figure 3-17: Speed, power, vacuum and discharge pressure of the main pump vs. time. ......................................94
Figure 3-18: Speed, power, vacuum and discharge pressure of the booster pump vs. time. ..................................94
Figure 3-19: Line speed, density, total power and situ production as a function of time, with flow control. ........95
Figure 3-20: Speed, power, vacuum and discharge pressure of the booster pump vs. time, with flow control......95
Figure 4-1: The Moody diagram determined with the Swamee Jain equation. ....................................................107
Figure 4-2: The Darcy-Weisbach friction factor λl for smooth pipes as a function of the line speed vls. .............109
Figure 4-3: The Darcy-Weisbach friction factor λl for smooth pipes as a function of the pipe diameter Dp. ......109
Figure 4-4: Collected relative viscosity data from 16 sources by Thomas (1965). ..............................................111
Figure 4-5: Collected relative viscosity data from 16 sources by Thomas (1965), reduced. ...............................111
Figure 4-6: Markham fines Boothroyde et al. (1979), without Thomas (1965) viscosity. ..................................112
Figure 4-7: Markham fines Boothroyde et al. (1979), with Thomas (1965) viscosity. .......................................112
Figure 4-8: Iron ore Thomas (1976), without Thomas (1965) viscosity. .............................................................113
Figure 4-9: Iron ore Thomas (1976), with Thomas (1965) viscosity. ..................................................................113
Figure 5-1: The relative excess hydraulic gradient as a function of the hydraulic gradient, constant Cvs and
Dp=0.1524 m. ................................................................................................................................122
Figure 5-2: The relative excess hydraulic gradient as a function of the hydraulic gradient, constant Cvt and
Dp=0.1524 m. ................................................................................................................................122
Figure 5-3: The relative excess hydraulic gradient as a function of the hydraulic gradient, constant C vs and Dp=1
m. ..................................................................................................................................................123
Figure 5-4: The relative excess hydraulic gradient as a function of the hydraulic gradient, constant Cvt and Dp=1
m. ..................................................................................................................................................123
Figure 5-5: The hydraulic gradient im, il and excess hydraulic gradient im-il. Constant spatial volumetric
concentration. ................................................................................................................................126
Figure 5-6: The hydraulic gradient im, il and excess hydraulic gradient im-il. Constant delivered (transport)
volumetric concentration. ..............................................................................................................126
Figure 5-7: Behavior of narrow graded crushed granite slurry after Clift et al. (1982). ......................................127
Figure 5-8: Behavior of narrow graded crushed granite slurry after Clift et al. (1982). ......................................127
Figure 5-9: The 3 main flow regimes for fine particles. .......................................................................................130
Figure 5-10: The 4 main flow regimes for medium particles. ..............................................................................130
Figure 5-11: The 4 main flow regimes for coarse particles. .................................................................................131
Figure 5-12: The 3 main flow regimes for very coarse particles, including sliding flow. ....................................131
Figure 5-13: An example of a flow regime diagram. ...........................................................................................133
Figure 5-14: The definitions for fully stratified flow. ..........................................................................................137
Figure 5-15: The definition of the pressure losses, scenario’s L1 and R1, E rhg(il). ..............................................143
Figure 5-16: The definition of the pressure losses, scenario’s L1 and R1, i m(vls). ...............................................144
Figure 5-17: The definition of the pressure losses, scenario’s L2 and R2, E rhg(il). ..............................................145
Figure 5-18: The definition of the pressure losses, scenario’s L2 and R2, im(vls). ...............................................146
Figure 5-19: The definition of the pressure losses, scenario’s L3 and R3, E rhg(il). ..............................................147
Figure 5-20: The definition of the pressure losses, scenario’s L3 and R3, i m(vls). ...............................................148
Figure 5-21: Kazanskij (1980), sand, low concentration......................................................................................150
Figure 5-22: Kazanskij (1980), sand, high concentration ....................................................................................150
Figure 5-23: Clift et al. (1982), narrow graded crushed granite. ..........................................................................151
Figure 5-24: Clift et al. (1982), broad graded crushed granite. ............................................................................151
Figure 5-25: Wiedenroth (1967), coarse sand. .....................................................................................................152
Figure 5-26: Wiedenroth (1967), medium sand. ..................................................................................................152
Figure 5-27: Newitt et al. (1955), MnO2. ............................................................................................................153
Figure 5-28: Doron & Barnea (1993), Acetal. .....................................................................................................153
Figure 5-29: Babcock (1970), sand. .....................................................................................................................154

Page 300 of 304 TOC Copyright © Dr.ir. S.A. Miedema


List of Figures.

Figure 5-30: Thomas (1976), iron ore. .................................................................................................................154


Figure 5-31: Boothroyde (1979), gravel...............................................................................................................155
Figure 5-32: Wiedenroth (1967), gravel...............................................................................................................155
Figure 6-1: The Erhg parameter versus the relative line speed. .............................................................................165
Figure 6-2: The algorithm to determine the Limit Deposit Velocity....................................................................168
Figure 6-3: The resulting FLcurves.......................................................................................................................171
Figure 6-4: The Jufin-Lopatin ψ* compared to Gibert and DHLLDV. ................................................................176
Figure 6-5: The reciprocal particle Froude number of Jufin-Lopatin, Gibert and DHLLDV. .............................177
Figure 6-6: The power in the Wilson et al. (1992) model, d=0.2 mm. .................................................................186
Figure 6-7: The power in the Wilson et al. (1992) model, d=2 mm. ....................................................................186
Figure 6-8: The power in the Wilson et al. (1992) model, d=20 mm. ..................................................................187
Figure 6-9: The data of Durand & Condolios (1952). ..........................................................................................188
Figure 6-10: The v50 curves for a Dp=0.1524 m pipe. ..........................................................................................189
Figure 6-11: The v50 curves for a Dp=0.4 m pipe. ................................................................................................190
Figure 6-12: The v50 curves for a Dp=1.2 m pipe. ................................................................................................191
Figure 6-13: The data of Whithlock et al. (2004).................................................................................................193
Figure 6-14: The data of Blythe & Czarnotta (1995). ..........................................................................................193
Figure 6-15: The demi-McDonald of Wilson (1979). ..........................................................................................196
Figure 7-1: The display of the top view of the cutterdredge, also showing the channel. .....................................207
Figure 7-2: The display of the back view of the cutterdredge, also showing the cross-sectional channel profile.
.......................................................................................................................................................208
Figure 7-3: The display of the side view of the cutterdredge, also showing the longitudinal channel profile. ...209
Figure 7-4: The output of the winch parameters. .................................................................................................210
Figure 7-5: The coordinate system with the dredge in the neutral position. ........................................................210
Figure 7-6: The coordinate system with the dredge at a swing angle s. .............................................................211
Figure 7-7: The torque-speed characteristic of the winches. ................................................................................212
Figure 7-8: The torque-speed characteristic of the winches with the setpoints. Case where the required torque is
sufficient........................................................................................................................................213
Figure 7-9: The torque-speed characteristic of the winches with the setpoints. Case where the required torque in
the setpoint is not sufficient. .........................................................................................................213
Figure 7-10: The torque-speed characteristic of the winches with the setpoints. Case where the setpoint is smaller
then the actual revolutions. ............................................................................................................213
Figure 7-11: The dredge, winch and channel layout. ...........................................................................................214
Figure 7-12: The dredge and anchor layout for case 1, port. ...............................................................................214
Figure 7-13: The dredge and anchor layout for case 1, starboard. ......................................................................214
Figure 7-14: The rope speeds and forces for case 1. ............................................................................................216
Figure 7-15: The rope speeds and forces for case 2. ...........................................................................................216
Figure 7-16: The dredge and anchor layout for case 2, port. ...............................................................................217
Figure 7-17: The dredge and anchor layout for case 2, starboard. ......................................................................217
Figure 8-1: Phase 1 of the loading cycle. ..........................................................................................................225
Figure 8-2: Phase 2 of the loading cycle. ..........................................................................................................225
Figure 8-3: Phase 3 of the loading cycle. ..........................................................................................................226
Figure 8-4: Phase 4 of the loading cycle. ..........................................................................................................226
Figure 8-5: Phase 5 of the loading cycle. ..........................................................................................................226
Figure 8-6: Phase 6 of the loading cycle. ..........................................................................................................226
Figure 8-7: Phase 7 of the loading cycle. ..........................................................................................................227
Figure 8-8: Phase 8 of the loading cycle. ..........................................................................................................227
Figure 8-9: The loading cycle of a TSHD. ........................................................................................................228
Figure 8-10: The loading part of the cycle of a TSHD. ...................................................................................228
Figure 8-11: A sharp crested weir. .......................................................................................................................232
Figure 8-12: Values for the coefficient Ce as a function of ha/hb=h/M.................................................................232
Figure 8-13: An example of a loading cycle of a TSHD with many turns. ..........................................................233
Figure 8-14: A close up of the hopper volume registration. .................................................................................234
Figure 8-15: The layer thickness during a turn, registration and approximation. ................................................234
Figure 8-16: The cycle as registered is simulated with the theoretical model. .....................................................235
Figure 8-17: The decreasing of the height of the layer of water above the overflow at the end of the cycle. ......235
Figure 8-18: Loading curves according to Miedema & van Rhee (2007) with and without time delay. .............236
Figure 8-19: The top view of the ideal basin. .......................................................................................................239
Figure 8-20: The side view of the ideal basin. .....................................................................................................239
Figure 8-21: The path of a particle with a settling velocity greater than the hopper load parameter. ..................240

Copyright © Dr.ir. S.A. Miedema TOC Page 301 of 304


Introduction Dredging Engineering.

Figure 8-22: The path of a particle with a settling velocity equal to the hopper load parameter. ........................240
Figure 8-23: The path of a particle with a settling velocity smaller than the hopper load parameter...................240
Figure 8-24: The path of a particle with a non-uniform velocity distribution. .....................................................241
Figure 8-25: The effect of a rising sediment level. ..............................................................................................242
Figure 8-26: Determination of the basin settling efficiency. ................................................................................243
Figure 8-27: A graphical method to determine the settling efficiency. ................................................................243
Figure 8-28: The equilibrium of forces on a particle. ..........................................................................................244
Figure 8-29: The total settling efficiency for λ=0.01. ..........................................................................................246
Figure 8-30: The total settling efficiency for λ=0.02. ..........................................................................................247
Figure 8-31: The total settling efficiency for λ=0.03. ..........................................................................................247
Figure 8-32: The 0.4 mm grain distribution. ........................................................................................................250
Figure 8-33: The loading curves of the Small TSHD. ..........................................................................................250
Figure 8-34: The loading curves of the Jumbo TSHD. ........................................................................................251
Figure 8-35: The loading curves of the Mega TSHD. ..........................................................................................251
Figure 8-36: Overview of the 2DV model. ..........................................................................................................252
Figure 8-37: Loaded TDS and overflow losses as a function of time for a Small size TSHD. ............................254
Figure 8-38: Loaded TDS and overflow losses as a function of time for Jumbo TSHD. ....................................254
Figure 8-39: Loaded TDS and overflow losses as a function of time for a mega TSHD. ...................................255
Figure 8-40: Comparison of the two models for the Small hopper. .....................................................................256
Figure 8-41: Comparison of the two models for the Jumbo hopper. ....................................................................257
Figure 8-42: Comparison of the two models for the Mega hopper. .....................................................................257
Figure 8-43: The 4 grain distributions. .................................................................................................................261
Figure 8-44: The loading curves for the Small TSHD. ........................................................................................265
Figure 8-45: The loading curves for the Mega TSHD. ........................................................................................265
Figure 8-46: The loading curves including the storage effect for the Small TSHD. ............................................266
Figure 8-47: The loading curves including the storage effect for the Mega TSHD. ............................................266
Figure 8-48: The grain distribution curves, original, overflow losses and sediment for the Small TSHD. .........267
Figure 8-49: The grain distribution curves, original, overflow losses and sediment for the Mega TSHD. ..........267
Figure 8-50: The overflow losses compared with an analytical model for the Small TSHD. ..............................268
Figure 8-51: The overflow losses compared with an analytical model for the Mega TSHD. ..............................269
Figure 8-52: The 7 levels of erosion according to Delft Hydraulics (1972).........................................................276

Page 302 of 304 TOC Copyright © Dr.ir. S.A. Miedema


List of Tables.

Chapter 11: List of Tables.


Table 1-1: Soil Classification. ................................................................................................................................. 5
Table 1-2: Empirical values for ρt, of granular soils based on the standard penetration number, (from Bowels,
Foundation Analysis). .....................................................................................................................21
Table 1-3: Empirical values for ρs, of cohesive soils based on the standard penetration number, (From Bowels,
Foundation Analysis). .....................................................................................................................21
Table 1-4: Typical Soil Characteristics (From Lindeburg, Civil Engineering Reference Manual for the PE Exam,
8th edition). ......................................................................................................................................21
Table 1-5: Typical Values of Soil Index Properties ...............................................................................................21
Table 1-6: Designation of Granular Soil Based on Relative Density. ....................................................................22
Table 1-7: Empirical values for φ, of granular soils based on the standard penetration number, (From Bowels,
Foundation Analysis). .....................................................................................................................26
Table 1-8: Relationship between φ, and standard penetration number for sands, ..................................................26
Table 1-9: Relationship between φ, and standard penetration number for sands, ..................................................26
Table 1-10: External friction angle φ values. .........................................................................................................27
Table 1-11: Guide for Consistency of Fine-Grained Soil, NAVFAC 7.02 ............................................................29
Table 1-12: Empirical Values for Consistency of Cohesive Soil, (from Foundation Analysis, Bowels) ...............29
Table 1-13: The Mohs scale (source Wikipedia). ..................................................................................................32
Table 1-14: Typical values of the permeability k. ..................................................................................................37
Table 1-15: Some permeabilities according to Hazen’s equation. .........................................................................37
Table 1.2-1: The 8 possible flow regimes. ...........................................................................................................138
Table 1.2-2: Scenario’s L1 and R1.......................................................................................................................143
Table 1.2-3: Indication of occurrence of L1.........................................................................................................143
Table 1.2-4: Indication of occurrence of R1. .......................................................................................................144
Table 1.2-5: Scenario’s L2 and R2.......................................................................................................................145
Table 1.2-6: Indication of occurrence of L2. ........................................................................................................145
Table 1.2-7: Indication of occurrence of R2. .......................................................................................................146
Table 1.2-8: Scenario’s L3 and R3.......................................................................................................................147
Table 1.2-9: Indication of occurrence of L3. ........................................................................................................147
Table 1.2-10: Indication of occurrence of R3. .....................................................................................................148
Table 6-1: Some A and ACv values. ....................................................................................................................167
Table 6-2: Group classification of Jufin-Lopatin (1966), source Kazanskij (1972). ............................................175
Table 6-3: Correction factor a, source Kazanskij (1972) . ...................................................................................179
Table 8-1: The data of the TSHD used. ...............................................................................................................236
Table 8-2: The data of the TSHD's used. .............................................................................................................249
Table 8-3: The hopper content after the filling phase. .........................................................................................249
Table 8-4: The main dimensions of the 4 TSHD's. ..............................................................................................260
Table 8-5: Additional and derived quantities. ......................................................................................................260
Table 8-6: The characteristics of the 4 grain distributions. ..................................................................................260
Table 8-7: The simulation results with the 0.400 mm sand..................................................................................261
Table 8-8: The simulation results with the 0.250 mm sand..................................................................................261
Table 8-9: The simulation results with the 0.150 mm sand..................................................................................261
Table 8-10: The simulation results with the 0.100 mm sand................................................................................262

Copyright © Dr.ir. S.A. Miedema TOC Page 303 of 304


Introduction Dredging Engineering.

Page 304 of 304 TOC Copyright © Dr.ir. S.A. Miedema


Introduction Dredging Engineering.

OE4607 Introduction Dredging Engineering.


MSc Offshore & Dredging Engineering
Delft University of Technology
2nd Edition

by
Dr.ir. Sape A. Miedema

Page 2 of 2 TOC Copyright © Dr.ir. S.A. Miedema

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy