Slurry Transport Using Centrifugal Pumps (4th Edition) 2023
Slurry Transport Using Centrifugal Pumps (4th Edition) 2023
Václav Matoušek
Lionel Pullum
Anders Sellgren
Slurry Transport
Using Centrifugal
Pumps
Fourth Edition
Slurry Transport Using Centrifugal Pumps
Robert Visintainer • Václav Matoušek
Lionel Pullum • Anders Sellgren
© The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature Switzerland
AG 1996, 2006, 2023
This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher, whether
the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of
illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and
transmission or information storage and retrieval, electronic adaptation, computer software, or by
similar or dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors, and the editors are safe to assume that the advice and information in this
book are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or
the editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.
This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
To the memory of Ken Wilson and Graeme
Addie, whose insight, energy, dedication, and
imagination have left an enduring mark on the
field of slurry transport.
Foreword by Robert Cooke
This important engineering textbook is a legacy of Georgia Iron Works’ long history
of slurry pump and pipeline testing, which started in 1956 and led to the formation of
the GIW Industries Hydraulic Laboratory in Grovetown. The Laboratory established
a tradition of diligently and accurately performing test investigations to gain insights
into slurry behavior and equipment performance which continues today.
Early pioneers used data generated by the Laboratory to support the development
of modelling approaches which are the foundation of our modern day understanding
of slurry flow behavior. These pioneers included Ken Wilson, Graeme Addie, and
Roland Clift, who in 1992 published the first edition of Slurry Transport Using
Centrifugal Pumps. This landmark publication provided engineers with the first
comprehensive reference on the design and operation of slurry pump and pipeline
transport systems.
The importance of this book is demonstrated by continually remaining in print,
and now, 30 years later, publication of this fourth edition of Slurry Transport Using
Centrifugal Pumps, authored by a new generation of contributors, which updates the
original text with new insights and understanding, and improved slurry analysis
methodologies.
My first exposure to the GIW Hydraulic Laboratory was indirect. My professor
spent a sabbatical with the Laboratory in the early 1980s and brought back GIW’s
methodology to establish a slurry laboratory at the University of Cape Town. As a
result, my initial training in slurry transport reflected the Laboratory’s approach and
philosophy. Many years later, I am honored to be included as an instructor for the
annual “Transportation of Solids using Centrifugal Pumps” short course sponsored
by the Hydraulic Laboratory.
Slurry Transport Using Centrifugal Pumps should be in the reference library of
every engineer involved in the design and operation of slurry systems.
vii
Foreword by Cees van Rhee
ix
x Foreword by Cees van Rhee
I am sure that this fourth edition will be the basic textbook for academics and
engineers active in the field of pumps and slurry transport for years to come.
Much of the material presented in this book has been developed and compiled in
connection with a short course presented annually since 1978 under the sponsorship
of the GIW Industries Hydraulic Lab (Figs. 1 and 2), and known throughout the
industry as the “GIW Slurry Course.” Earlier editions of the book were authored by
the founders and primary instructors of this course, who were also regular contrib-
utors to the body of research and scientific literature on these topics. The authors of
the current edition have continued this tradition of serving both course and book,
regularly teaching and refining the concepts covered within these pages for the
benefit of each new generation of engineers.
This book is divided into two main parts. The first half, up to the end of Chap. 7, is
concerned primarily with the classification and behavior of the most common slurry
types. The individual properties of the solids and liquids which comprise a slurry are
considered with respect to their influence on key properties of slurry flow affecting
system design, such as the onset of solids deposition and the pressure gradient within
a pipeline. Both general guidance and specific models are covered. The second
portion of the book is concerned with the design and behavior of centrifugal pumps
handling slurries, and the way in which the pumps and pipelines interact as a system.
Within the first half of the book, Chap. 2 leads into the work’s main topics by
providing a review of the basic principles of fluid mechanics in pipeline flow and the
settling of solid particles in liquids, both of which form a necessary background for
slurry flow analysis. This is needed as a ready reference for points on which some
readers may wish to “brush up” and to establish a common basis for those of
different scientific and engineering backgrounds.
Chapter 3 provides a system of classification for the slurry types covered in the
remainder of the book. These are broadly divided into settling and non-settling
slurries, with subtypes and exceptions explained in overview along with their key
behavioral aspects.
xi
xii Preface
Fig. 1 The GIW Industries Hydraulic Lab with pump and sump setup for a 200 mm slurry pipeline
test
Fig. 2 Exterior view of GIW Hydraulic Lab showing pipeline test loops
A number of worked examples are included throughout the text in order to clarify
the ideas presented and give examples of how they are used in slurry system design.
In general, SI units are used throughout, with reference to customary US units only
when these are in common international usage.
This fourth edition has been largely rewritten with the needs of the practicing
engineer in mind, including slurry system designers, operators, and reliability
engineers, while at the same time providing an introduction to the topic suitable
for the upper level engineering student.
The subject matter has been reordered according to the practical development of
the methods presented, and a new chapter added on the principles and classifications
of slurry flow. Care has been taken to formalize the theoretical concepts, providing
greater clarity of the slurry flow dynamics upon which the various models are based.
In the realm of settling slurries, the 4-Component Models for pipeline flow and
pump solids effect have been updated according to an extensive series of full-sized
tests. The treatment of complex slurries has also been expanded to include a broader
discussion of non-Newtonian fluids and their interaction with coarse particles.
A new chapter on testing methods and instrumentation has been added, which
provides a summary of the instrumentation commonly used in slurry systems, and
presents strategies and rheological techniques suitable for characterizing slurry
flows, along with much practical advice.
The sections on erosive wear and operational experience have been updated,
including a new and comprehensive parametric study of slurry pump wear and
operating cost, and an expanded analysis of system stability and operability.
This book is the fourth edition of a text that emerged from an annual course on slurry
transport, hosted by the GIW Industries Hydraulic Lab over the past 45 years in
Augusta and Grovetown, Georgia, USA. The authors wish to acknowledge those
who have supported the “GIW Slurry Course” and have directly or indirectly
contributed to this book.
We owe a particular debt of gratitude to the founders of both course and book:
Dr. Kenneth Wilson, Mr. Graeme Addie, and Dr. Roland Clift, who formed the
primary teaching staff for the course during its first 30 years, and who co-authored
the previous three editions. Their dedicated and holistic approach to advancement in
the field of slurry transport—experimental, theoretical, and practical—and their
large body of contributions, which go well beyond the contents of this book, form
an enduring legacy that deserves special mention.
We also wish to acknowledge the contributions and participation of Dr. M.-
R. Carstens, Mr. Reab Berry, Dr. R. Alan Duckworth, Dr. M. Roco, and Mr. Lee
Whitlock, who supported the course in past years. Particular recognition is due to
Mr. Thomas W. Hagler, Jr., former CEO of GIW Industries, who, along with
Graeme Addie, provided the vision and practical support needed to develop the
GIW Hydraulic Lab in its early years.
For their contributions to this edition and their ongoing support of the course, we
thank Dr. Krishnan Pagalthivarthi, Dr. Harry Tian, Mr. John Harding, Mr. Brian
Prochaska, Dr. Robert Cooke, and Dr. Mohamed Garman, as well as the members of
the GIW staff who supported the production of this edition, including Mrs. Janina
Baxley Rogers, Ms. Jennifer Belgin, and Mrs. Eileen Eubanks. Special thanks are
due to the staff of the GIW Hydraulic Lab, led for the past ten years by Mr. George
McCall. Without their hard work and dedication, the course as we know it today
could not exist. We also acknowledge the contributions of the late Dr. John Furlan,
who assisted the authors in reviewing the first drafts of the text, and whose enthu-
siastic participation in the course will be remembered by all who met him.
xv
xvi Acknowledgments
Finally, we wish to express our thanks to the current leadership of GIW Industries
and its parent company, KSB SE & Co. KGaA, including Mr. Wolfgang Demmler
and Mr. Jonathan Samuel, who have continued to support the vision of excellence in
research and education for which the book, the short course, and the GIW Hydraulic
Lab have become known.
Contents
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Applications of Slurry Transport . . . . . . . . . . . . . . . . . . . . . . . 1
1.1.1 Metals and Minerals . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.1.2 Dredging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.1.3 Specialty Applications . . . . . . . . . . . . . . . . . . . . . . . 6
1.2 The Blind Men and the Elephant . . . . . . . . . . . . . . . . . . . . . . 7
Reference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2 Review of Fluid and Particle Mechanics . . . . . . . . . . . . . . . . . . . . . 9
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2 Classification of Fluids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2.1 Rheological Properties . . . . . . . . . . . . . . . . . . . . . . . 12
2.2.2 Common Simple Non-Newtonian Fluids . . . . . . . . . . 14
2.3 Classification of Solids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.3.1 Solids Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.3.2 Particle Diameter . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.3.3 Particle Shape . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.4 Basic Relations for Fluid Flow in Pipelines . . . . . . . . . . . . . . . 23
2.4.1 Conservation of Continuity and Momentum . . . . . . . . 23
2.4.2 Bernoulli’s Equation: Head and Hydraulic
Gradients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.5 Pipeline Friction of Newtonian Fluids . . . . . . . . . . . . . . . . . . . 29
2.5.1 Example 2.1: Flow of Water in a Pipe-Pump
System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
2.6 Settling of Solids in Newtonian Fluids . . . . . . . . . . . . . . . . . . 43
2.6.1 Example 2.2: Calculation of Terminal
and Hindered Settling Velocities . . . . . . . . . . . . . . . . 51
2.7 Settling of Solids in Non-Newtonian Fluids . . . . . . . . . . . . . . . 52
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
xvii
xviii Contents
Symbols
The principal symbols used in this book are listed below, together with brief
definitions and the units most commonly used by practicing engineers. Dimension-
less quantities are indicated by (–).
Abbreviations
xxxiii
xxxiv About the Authors
Fig. 1.1 Mid-test examination of 50-mm granite being pumped in an 8-inch (200 mm) pipeline for
simulation of deep sea mining solids transport
The extraction of metal ores from the earth’s crust involves the movement of vast
tonnages of solids since the typical ore body contains only a few percent (and in
some cases much less) of the metal-bearing minerals. The mined solids must often be
moved multiple times: first into a processing plant, then through various steps to
separate the useful minerals, and finally, the remainders (or tailings) must be
returned to the environment in a sustainable way, often at some distance from the
plant. In many cases, the bulk of this processing and transport is done in slurry form.
By way of example, Fig. 1.2 shows a typical “hard rock” mining circuit, as might
be used in a copper mine. Ore is brought to the plant by truck or conveyor, where it is
mixed with water and loaded into grinding mills. From this point onwards, it remains
in slurry form, exiting the mills as a mixture of water, coarse gravel, and fine sand.
Recirculation pumps feed this slurry through cyclone separators which divide it into
1.1 Applications of Slurry Transport 3
Fig. 1.2 A typical hard rock mining circuit: (a) grinding mills, (b) recirculation pump, (c) cyclone
separators, (d) froth and concentrate handling pumps, (e) flotation cells, (f) tailings thickener, (g)
tailings pumps, (h) tailings pipeline. (Courtesy GIW Industries, Inc)
two streams according to particle size, the coarse solids being returned to the mills
for further grinding, and the fine sand slurry piped to floatation cells where the
valuable mineral or concentrate is separated from the remaining tailings. These two
streams (concentrate and tailings) continue in slurry form, the former to be processed
for shipment and the latter to be concentrated or thickened and again pumped to a
tailings storage area, which may be kilometers distant, or perhaps back into vacated
parts of the mine itself. During this process, the slurry streams change considerably
in their character and appearance due to differences in the size, density, and
concentration of the entrained solids and the temperature, viscosity, and chemical
composition of the liquid mixtures which carry them. The centrifugal pumps used in
hard rock mining can vary from 20 kW machines with impeller diameters of 200 mm
or less operating in 50-mm (2-inch) pipelines to 4000-kW machines with impeller
diameters of 2 m or more operating in 750-mm (30-inch) pipelines.
The above example is simplified, and complete volumes could be (and in fact
already have been) written by others about the many processes used to separate and
concentrate metal ores. Some of the more reactive metals, such as aluminum and
lithium, require complex chemical processes to liberate the metals; and again, slurry
handling plays a key role. The important fact is that many—perhaps most—of these
processes contain multiple slurry streams with centrifugal pumps and pipelines, the
character of which vary almost as much as the processes themselves.
Other minerals are commonly mined and transported as slurries using centrifugal
pumps, including phosphate rock for fertilizer production and oil sands for bitumen
extraction. In these cases, the mined solids may be transported long distances from
the extraction site to the processing plant, requiring large pumps of high-pressure
design (Fig. 1.3). In the case of the Florida phosphate mines in the United States, an
as-mined mixture or “matrix” of phosphate rock, gravel, and clay is pumped as far as
27 km to the processing plant in 500-mm (20-inch) pipelines using 1500-kW skid-
4 1 Introduction
Fig. 1.3 High-pressure centrifugal pump for long-distance slurry transport (GIW TBC-92)
mounted centrifugal pumps. The first pump is typically outfitted with a flexible
suction line to pull the matrix from a temporary pit constructed at the mine site for
mixing the matrix into a slurry (Fig. 1.4). Additional pumps are spaced along the line
singly or in pairs, with up to 20 pumps used on the longest lines. In oil sand mining,
the ore is actively processed as it moves through the pipeline, changing its properties
along the way, so that the pumps and pipeline serve as both transport system and
process equipment.
Some metals and minerals occur in alluvial deposits washed down from eroded
ore bodies, including titanium, tungsten, placer gold, and diamonds. These may be
found in ancient sand deposits, active rivers, or even offshore in several hundred
meters of seawater. Large quantities of silica, one of the most abundant minerals on
earth, are still mined today for the sand and gravel used in the construction of
buildings, foundations, and roads and for use in filtration, metal casting, fracking
proppants, and the manufacture of glass. Slurry transport is widely used in mining,
conveying, washing, and sizing these materials. Clays are also mined by slurry
processes for a variety of purposes, both industrial and commercial. The washing
and transport of coal for power generation, although in decline, remains an important
application.
1.1 Applications of Slurry Transport 5
1.1.2 Dredging
Another important use of slurry transport, representing perhaps the largest move-
ment of solids anywhere on earth, is found in dredging applications. These include
waterway depth maintenance for navigation, coastline modification for beach pro-
tection, environmental remediation, and port construction. Land building, involving
the moving of offshore deposits to create new land, is carried out for many reasons,
from port construction to artificial islands, for coastline protection, transportation
infrastructure, and recreation. Once again, there is a broad range in the scale of
operations and the slurry properties encountered in these applications, from small
projects in lakes and streams handling a few cubic meters per hour to massive trailing
suction hopper dredges operating in the open ocean and handling multiple cubic
meters per second (Fig. 1.5). Pipelines can exceed 1 m in diameter and the pumps
may operate at power levels beyond 10 MW (Fig. 1.6).
The slurries themselves also vary, although perhaps not so widely as seen in the
mineral processing industry. Solids range from micron-sized clay particles through
muds and silts, fine and coarse sands, gravel, shell, and rocks up to 550 mm in
diameter. Dredging operations are also faced with handling unexpected debris
present in the environment, from tree roots to unexploded ordinance.
More recently, attempts are being made to recover high-quality metal ores that
have been deposited by geothermal processes or otherwise precipitated at depths of
4–5 km in the open ocean. At the time of writing, the technology to economically
and sustainably mine and transport these ores to the surface is still being developed.
6 1 Introduction
Fig. 1.5 Trailing hopper suction dredge with “rainbow” discharge for land-building operations
Fig. 1.6 A 1100-mm discharge, 10-MW double wall dredge pump (GIW DWD-2500)
Due to the diversity of applications, slurry behavior varies widely. Some closely
approximate pure fluids, while others behave in an almost solid fashion. The
complex mixtures in between these extremes may have unexpected and sometimes
counterintuitive properties. The engineers dealing with these flows also come from a
variety of backgrounds, including civil, mechanical, mineral, and chemical, and have
developed different experiences with the design and operation of slurry systems.
This diversity in slurry behavior and practical experience has been noted since the
earliest editions of this book and the short course that it supports (see the Preface).
Although the technical content of both the book and course has advanced consider-
ably over the years, one point has remained constant: the invocation of the old tale of
the blind men and the elephant (Saxe 1872). In this fable, six blind men who had
always wanted to know what elephants were like were finally allowed to spend a few
minutes in contact with one of these beasts. But since each man touched a different
part of the animal, they later found themselves in greater disagreement than before
on the nature of the elephant, comparing it to everything from a house (based on its
side) to a rope (based on its tail) (Fig. 1.7).
As we begin our journey through the topic of slurry transport, we will endeavor to
bring the entire elephant into focus in all of its majestic glory, from the theoretical
aspects of its inner workings to the practical considerations of its daily care and
feeding. For those of our readers who are responsible for managing such elephants,
we hope you find these pages both interesting and instructive.
For a detailed overview of the topics covered in this book, as well as the particular
updates made in the fourth edition, refer to the Preface.
Reference
J.G. Saxe, The blind men and the elephant. Fables and legends of many countries rendered in
rhyme, in The Home Book of Verse, ed. by B.E. Stevenson, (Holt, Reinhart and Winston,
New York, 1872), pp. 1877–1879
Chapter 2
Review of Fluid and Particle Mechanics
2.1 Introduction
Before considering the flow of mixtures containing liquid and solid particles that
behave as separate phases, we must first examine the flow of liquids and the motion
of particles in liquids. In discussing these topics, we will also introduce some of the
concepts, terminology, and notation used in later chapters. This chapter is intended
as a review primarily to reinforce knowledge that readers will have encountered in
their undergraduate engineering curriculum, but may not have used in the interim.
As in other parts of this book, the level of presentation is directed to practical
engineering application. This chapter is not intended as an introduction to fluid
mechanics in general, to turbulence, or to the micromechanics of particle-fluid
systems. While those subjects have significance for fundamental research, they are
not required for an engineering treatment of slurry flow.
To characterize a simple fluid such as water, only two material properties are
required: density and viscosity. Density, denoted by ρ, represents the mass of fluid
per unit volume. Viscosity is a measure of the fluid’s resistance to deformation by
shearing.
Viscosity is best illustrated using the conceptual experiment shown in Fig. 2.1, in
which a fluid fills a gap of thickness y between two flat parallel plates. The plate
forming the y = 0 plane is kept stationary, while the other plate (at y = Y ) is moved
parallel to the first at a steady velocity (U ). The fluid immediately in contact with
each solid plate keeps the same velocity as the plate; this phenomenon is known as
the non-slip boundary condition. Thus, the velocity of the fluid at y = 0 is zero, and
the velocity at y = Y is U. At any intermediate position, the fluid velocity, u, is yU/Y.
y y
U t
y=Y
y=0
U=0 t U u
1.0
The quantity U/Y is the velocity gradient in the fluid, denoted by du/dy. Its signif-
icance is that it represents the rate of shear deformation of the fluid. It is therefore
known as the rate of shear strain, strain rate, or, more simply, the shear rate.
To maintain the steady motion in Fig. 2.1, it is necessary to apply a force to the
moving plate and an equal and opposite restraining force to the stationary plate.
These forces are parallel to the plates and are in the direction of their relative motion.
The force on one plate per unit area is known as the shear stress, denoted by τ. As
shown in Fig. 2.1, τ represents the shear stress exerted by each plate on the fluid in
the gap. Simple Newtonian mechanics dictates that the fluid exerts an equal and
opposite shear stress on the plate and that the shear stress within the fluid at any plane
parallel to the plates is also τ.
The viscosity of the fluid is defined as the ratio of the shear stress (τ) to the shear
rate (du/dy), and we could construct an experiment like that shown in Fig. 2.1 to
measure it. However, while convenient for a textbook explanation, the flat plate
experiment is inconvenient to execute. Numerous, more practical ways of measuring
viscosity are commonly used, and some of these are introduced below.
For many simple fluids, like water, the shear stress is linearly proportional to the
shear rate, as seen in Fig. 2.2.
2.2 Classification of Fluids 11
du
τ=μ ð2:1Þ
dy
Equation (2.1) is the constitutive equation of a Newtonian fluid. The ratio μ/ρ is
known as the kinematic viscosity of the fluid and is denoted by ν.
For practical purposes, liquids are incompressible over the range of conditions
encountered in industries using hydraulic transport; therefore, ρ and μ can be taken
as independent of pressure. Both ρ and μ do, however, depend on temperature, both
often (though not always) decreasing with increasing temperature. Table 2.1 pro-
vides values for the density, viscosity, and vapor pressure of water as a function of
temperature.
Close approximations of these values from 1 °C < T < 99 °C may be obtained
using Eqs. (2.2, 2.3, and 2.4) (Tilton and Taylor 1937; Korson et al. 1969; Lide
2005; McCutcheon et al. 1993).
For the density (kg/m3) of water at atmospheric pressure:
1
For general fluids, the dynamic viscosity will be denoted by the symbol η in this book, reserving
the use of μ specifically for Newtonian fluids.
12 2 Review of Fluid and Particle Mechanics
ðT - 3:9863Þ2 ðT þ 288:9414Þ
ρw ðT Þ = 1000 1- ð2:2Þ
508929:2ðT þ 68:12963Þ
For the vapor pressure (mm Hg) of water (from the Antoine equation):
1730:63
logðpv Þ = 8:07131 - ð2:4Þ
233:426 þ T
Figure 2.2 shows the linear relationship between applied shear stress and resulting
shear rate for a Newtonian fluid undergoing simple shear. While many common
fluids, such as air and water, behave in this Newtonian fashion, many other fluids do
not. Those of particular interest to slurry transport are high concentration, fine
particle slurries, and slurry mixtures containing significant amounts of clay which
behave, in many respects, like single-phase fluids, even though they contain both
solid and liquid components. When classifying these non-Newtonian fluids or
slurries, it is necessary to define the shear stress response over a range of shear
rates. Such measurements are made using a rheometer or viscometer. See Chap. 12
for details of these devices and the considerations that must be taken into account
when using them. It may also be important to identify the range of shear rates
applicable to the flow being examined or modeled, and typical shear rates for
common industries’ slurry processes are given in Table 2.2.
The more complex behavior of non-Newtonian slurries is the result of three-
dimensional particulate structures forming throughout the fluid. These structures,
which are formed through complex surface chemical effects and the resulting inter-
particle forces, hold the structures together and allow them to resist external forces.
They are often very strongly influenced by particle size and concentration. When the
slurry is flowing, these structures deform and reorganize under the imposed strain in
a non-linear way relative to the shear rate, resulting in a non-linear (non-Newtonian)
relationship between the shear stress and the shear rate.
The amount of deformation and reorganization that occurs is a strong function of
the applied stress (or shear rate). If the applied stress to the slurry is sufficiently low,
the particulate structures within the fluid may be able to resist all movement, and the
slurry behaves as a virtual solid. Increasing the applied stress will eventually
overcome this resistance. At a value known as the shear yield stress, or simply
yield stress, the slurry will start to shear and flow. Relaxing the shear stress (or strain)
2.2 Classification of Fluids 13
on the slurry enables these structures to reorganize and reform. Once sheared, these
reformed structures may not be as strong as that of the slurry before it was sheared. In
some cases, the strength may recover if the slurry is allowed to rest over a long
period of time, sometimes a matter of weeks. In other cases, it may recover only
partially, or not at all, depending on the nature of the particle structures and their
interactions.
Such behavior, known as thixotropy, is generally a strong function of the level of
shearing and the length of time that the slurry has been sheared (i.e., the slurry’s
shear history). If a slurry is sheared at a constant rate, and for long enough, the
internal structure will stabilize, and the slurry will reach an equilibrium state. Note
that this rate is not a fixed value but will generally be a function of the shear history.
While thixotropic behavior can be encountered when pumping slurries (e.g., in the
suction piping from a thickener underflow or in the disposal of red mud), the high
shear rates associated with centrifugal pumps are generally sufficient to break any
fluid structure down to an equilibrium condition, although cases exist where multiple
passes are required.2 We can therefore usually consider the slurry discharging from
the pump to have constant rheological properties and not be thixotropic for the
duration of its journey through the pipe. This is the approach adopted in this book.
Slurries may respond elastically over very short time scales, but for most
pumping applications, these time scales are much shorter than the time scales of
interest and the slurries may be considered, for all practical purposes, to be inelastic.
Note that such elastic behavior should not be ignored for processes that involve
highly dynamic shear histories, such as clay coating in paper production. Elastic
behavior will not be considered further in this book.
The modeling of non-Newtonian fluid slurry flows will be discussed further in
Chap. 6, with emphasis on the practical methods used by the slurry system designer.
2
In these cases, a special recirculation loop may be installed at the first pump to provide extra
shearing before the slurry enters the main pipeline.
14 2 Review of Fluid and Particle Mechanics
A non-Newtonian fluid is any fluid that does not comply with the definition of a
Newtonian fluid given above (Fig. 2.2), meaning it exhibits a variable relationship
between shear rate and shear stress or is sensitive to its shear history or is elastic.
Since this book is not concerned with thixotropic or elastic fluids, our characteriza-
tion of a non-Newtonian fluid will focus on its variable relationship between shear
stress and shear rate. Typical rheograms (plots of shear stress or viscosity vs. shear
rate) are shown in Fig. 2.3 for some generic non-Newtonian fluid types. Note for a
general fluid the symbol η is used to denote dynamic viscosity, reserving the symbol
μ specifically for Newtonian fluids.
The constitutive equations for the fluids shown in Fig. 2.3 are given in Table 2.3,
where μ, k, n, η0, η1, ηB,τyB, τyC, and τy are model parameters, the latter three being
yield stresses that must be overcome before flow can occur.3
As described previously, the Newtonian fluid model is typical of water and air.
The power-law fluid model, typical of molten polymers, some hydrocarbons, and
an approximation to some low-concentration slurries, is a particularly simple
non-Newtonian model that assumes a simple power-law variation of viscosity with
shear rate. In reality, such fluids typically exhibit Newtonian plateaus at both low and
high shear rates, and a slightly more complex model, the Cross model, captures this
behavior quite well. However, reference to Table 2.2 shows that most bulk pipe and
pump flows are limited in shear rates to values between 0 and 1000 reciprocal
seconds, and for general engineering purposes, a lower limit of 0.1 reciprocal
seconds is adequate. Over this shear rate range, differences in the variation in
viscosity for the power and Cross model fluids are minimal, as seen in Fig. 2.3b,
and the simpler power model can be successfully used in pump and pipeline designs.
Yield visco-plastic fluids, represented by the Bingham, Casson, and Herschel-
Bulkley model fluids, are of the most interest in slurry flows, as these models typify
the behavior of moderate to high concentration mixtures of fine particles, especially
where clay-type particles are involved. The Bingham plastic model is particularly
simple, representing the constant slope characteristic of a Newtonian fluid, but with a
yield stress, and this model has been used successfully for many years, especially for
slurries with a high clay content. The Casson model is also useful where there is a
slight non-linearity in response and models chocolate and blood flows very well. The
3
Note that model parameters have been chosen for illustrative purposes in Fig. 2.3. For consistency,
these same parameters have been used in Chap. 6 to produce radial profiles, etc.
2.2 Classification of Fluids 15
12
(a)
dilatant fluid?
10
shear stress, t ( Pa)
4 h Newtonian model
Power law model
1 Cross model
2 Bingham plastic model
Casson model
Herschel-Bulkley model
0
0 200 400 600 800 1000
.
shear rate, g (1/s)
100000
Newtonian model
10000 = Power law model
̇ Cross model
1000 Bingham plastic model
Viscosity, h (Pa s)
Casson model
100
Herschel-Bulkley model
10 dilatant fluid?
1
0.1
0.01
0.001
(b)
0.0001
0.0001 0.001 0.01 0.1 1 10 100 1000 10000 100000
.
shear rate, g (1/s)
Fig. 2.3 Fluid rheograms for various generic fluid types: (a) variation of shear stress with shear
rate; (b) variation of viscosity with shear rate. The viscosity for the Cross model fluid at a shear rate
of 400 reciprocal seconds is shown in (a) and as an open symbol in (b). A dilatant fluid is also
indicated
Table 2.3 Simple constitutive relationships for the generic fluid types shown in Fig. 2.3 in terms of
shear stress (τ) and viscosity (η) as a function of shear rate (_γ )
Model τðγ_ Þ ðPaÞ ηðγ_ Þ (Pa s)
Newtonian τ = μ_γ μ
Ostwald de Waele (power law) τ = k γ_ n k γ_ n - 1
η0 - η1
Cross model η - η1 η1 þ
τ = ˙_γ η1 þ 0 m 1 þ ðλ_γ Þm
1 þ ðλ_γ Þ
Bingham plastic τ = τyB þ ηB γ_ τyB
þ ηB
γ_
1
ðk γ_ Þ =2 1 2
1
Casson model 1=2þ
τ1=2 = τyC . τyC 2 þ ðk γ_ Þ2
γ_
Herschel-Bulkley fluid τ = τy þ k γ_ n τy
þ k γ_ n - 1
γ_
16 2 Review of Fluid and Particle Mechanics
Fig. 2.4 Particle types and densities (a) bulk density of an assemblage of particles, (b) density of a
compact particle, and (c) skeletal density of a porous particle
Finally, since most mineral slurries are water-based, it is unsurprising to find that
the vapor pressure for these slurries corresponds to the fluid’s vapor pressure (in this
case, water) at the slurry temperature (Cooke 2007).
There are three types of density associated with particles: the bulk density, which
corresponds to an assemblage of loosely poured particles; the density, which is a
density of a single solid particle; and the skeletal density, which is the density of the
actual solid material of a porous particle, as shown in Fig. 2.4.
These three densities are all different and of increasing numeric values. For
example, particles containing 5% by volume voids, a loosely packed concentration
of 70% by volume, and an underlying material density of 2600 kg/m3 would have
the following densities when immersed in water: bulk density = 2064 kg/m3;
particle density = 2520 kg/m3 (including the water-filled pores); skeletal density
= 2600 kg/m3.
In hydraulic conveying, the density of the individual particles will govern the way
in which they move through the conveying fluid. If the particles contain pores that
are flooded with the surrounding fluid, then the particle density must include the
mass of the fluid filling the pores. Whether or not the pores are filled depends on the
surface tension (usually dependent on temperature), the surrounding static pressure,
and the pore geometry. Fortunately, most ore and rock particles conveyed in the
industry have very low porosities, so the skeletal density may usually be used as an
approximation to the particle density without any great error. Table 2.4 lists some
typical ore and particle densities.
Methods to determine the density of solids are readily found in standard mineral
and solids processing handbooks (e.g., Rao 2017; Wills and Finch 2015).
18 2 Review of Fluid and Particle Mechanics
In modeling slurries, we are often correlating the slurry behavior to some defined
mean particle diameter derived from some selected measurement method. Particle
diameter measurement presents a particular challenge, as most slurry particles are
not spherical, nor are the particles likely to be all of the same size. Any number of
definitions may be used to characterize the diameter of a non-spherical particle.
These include the diameter of an equivalent spherical particle of equal volume
inferred via optical methods, the smallest spacing between wires in a square mesh
through which the particle will pass, or a simple measurement spanning the largest
particle dimension. Unless specified, only two types of particle diameter are used in
this text: an area equivalent diameter (da), described in Sect. 2.6, and mass median
diameters based on sieving, as described below. These are the most commonly used
types throughout the mineral processing industry and scientific literature and will be,
for brevity, denoted simply as d.
The sieving method consists of passing a sample of particles through a nest of
wire mesh sieves of decreasing gap size between wires. The mass of particles trapped
on each sieve is recorded and used to describe the particle size distribution. Diam-
eters determined in this way are known as mass median diameters and may also
represent volume mean diameters if the density and basic shape of the various solids
are constant. The data are usually presented as a table or plot of the percentage of the
total sample mass passing through each sieve versus the gap size of each sieve. This
plot is called the particle size distribution or PSD, and some examples are shown in
Fig. 2.5a with normalized mass fraction shown in Figure 2.5b. Similar plots can be
obtained with other methods based on mass or volume. “Narrow” distributions are
defined as those where the bulk of particles fall within a small size range, whereas
“broad” distributions span as a larger range, and “bi-modal” distributions contain
two separate groups of similarly sized particles separated by a size range with
relatively fewer particles. These terms are all somewhat subjective but are useful
for relative categorizations when comparing slurries.
Sieving and laser diffraction are the most commonly used methods in industry
today, but many other methods are available and include sedimentation methods,
Coulter counters, and automated microscopy. In some cases, more than one method
will be used to determine the particle size distribution. For example, one may use
simple sieving for the larger particles and laser diffraction for the smaller particles
within a given sample. In this case, it is best to overlap measurements so that
adjustment, through a regression analysis based on an assumed shape factor, can
be performed. Another challenge arises when results from one method must be
converted to that of another (see, e.g., Austin and Shah 1983; Murray and Holtum
1996; Shillabeer et al. 1992).
Various probability distributions can be used to smooth real data, e.g., normal,
log normal, and Rosin-Rammler or Weibull, but discrete values picked from the
cummulative plot (Fig. 2.5a) are usually sufficient to describe particle size and the
effect of the width of the distribution in most computations. The most commonly
20 2 Review of Fluid and Particle Mechanics
100
90
80
cummulative passing (%)
70
(ii) (iv) (iii) (i)
60
50
40
30
20
10
(a)
0
0.01 0.10 1.00 10.00 100.00
particle diameter (mm)
1.2
1.0
normalized mass fractions (-)
0.6
0.4
0.2
(b)
0.0
0.01 0.10 1.00 10.00 100.00
particle diameter (mm)
Fig. 2.5 Particle size distributions: (a) cummulative passing versus sieve size for (i) a coarse
narrow distribution; (ii) a fine distribution; (iii) a broad distribution; and (iv) a bi-modal distribution;
(b) normalized mass fractions for the distributions shown in (a)
used are d50 (representing the particle size at which 50% of the particles (by weight)
are smaller) and d85 (representing the analogous 85% threshold). Alternatively,
percentage fractions (by weight) may be defined according to fixed particle diame-
ters relevant to a particular model of slurry flow. For example, the weight fraction of
particles smaller than 40 microns (40 μm) is often identified, as these have the largest
effect on the properties of the carrier fluid. Some common diameters and diameter
ratios are shown in Table 2.5.
2.3 Classification of Solids 21
Table 2.5 Representative Distribution d50 (mm) d85 (mm) (d85 / d50) (-)
diameters for the particle
(i) Narrow 30.0 36 1.2
distributions shown
in Fig. 2.5 (ii) Typical 1.7 3.8 2.3
(iii) Broad 7.4 22.5 5.0
(iv) Bi-modal 7.2 20.6 2.9
Particle shape affects both the particle’s drag coefficient and the way in which the
particle orients itself during transport or settling. Shape characterization is a complex
field, and many definitions of the shape factor exist (see, e.g., Allen 1975), but
simple estimates are adequate for hydraulic conveying calculations. One common
shape factor is the particle’s sphericity (ψ) (Wadell 1935), defined as
where Ap and Vp are the surface area and volume of the particle, respectively, and as
ψ approaches a value of 1.0, the particle becomes more spherical. Figure 2.6 shows a
graphical display of particle sphericity, and Table 2.6 lists some values of sphericity
for various particles.
Although determining the sphericity for irregular particles is very difficult,
Aschenbrenner (1956) showed that the true sphericity could be approximated from
a combination of the flatness ratio, e1 = b=t, and the elongation ratio, e2 = l=b, based on
the particle’s, length, breadth, and thickness, as shown in Fig. 2.7.
The approximation of the particle’s sphericity (ψ w) is then given by
22 2 Review of Fluid and Particle Mechanics
1
12:2 e1 e22 3
ψw = qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ð2:6Þ
1 þ e2 ð1 þ e1 Þ þ 6 1 þ e22 1 þ e21
π ð0:928dG Þ2
ψw = ð2:7Þ
0:995SAellip
pffiffiffiffiffiffi
where dG = 3 lbt, and the elliptic surface area is given by
2.4 Basic Relations for Fluid Flow in Pipelines 23
0 11λ
λ λ λ
ðlb Þ þ ð lt Þ þ ð bt Þ
SAellip = 4π @ A ð2:8Þ
3 - 0:0942 1 - 27lbt ðl þ b þ t Þ - 3
and λ = 1.5349.
Reviewing the sphericities listed in the second column of Table 2.6, we can see
that a nominal value of ψ w ≈ 0.8 could be used for most particles commonly pumped
if actual values are not known.
Much of this book is concerned with the steady motion of fluids, in which the mean
velocity at any point does not change with time (although it may vary with location).
Whether the flow is steady or unsteady, analysis is based on three fundamental laws:
the continuity balance (or conservation of matter); linear or angular momentum
balances (which amount to the application of Newton’s laws to fluids); and the
mechanical energy balance (which is essentially the first law of thermodynamics
applied to fluids). For simplicity, and without loss of generality, we will limit the
discussion in this section to the flow of Newtonian fluids. Some simple applications
of these laws to solid–liquid mixtures are also considered.
Q VA DA DB VB
24 2 Review of Fluid and Particle Mechanics
4Q
VA = ð2:9Þ
πD2A
where DA is the internal pipe diameter at A, so that the pipe’s cross-sectional area is
πD2A =4. Similarly, at section B, the mean velocity is
4Q
VB = ð2:10Þ
πD2B
The equation of continuity for this section of pipe then takes the form:
πD2A V A πD2B V B
Q= = ð2:11Þ
4 4
For cases in which it is necessary to consider the local velocity in the pipe at
distance r from the axis (i.e., u(r)), the total volumetric flow is evaluated from the
integral:
ZD=2 ZD=2
Q= u2πrdr = 2π urdr ð2:12Þ
0 0
in which the area of the element from r to (r + dr) is 2πrdr and the velocity through it
is u. Specific applications of Eq. (2.12) are considered below, and in Chap. 3.
We next turn to the linear momentum equation. Consider the simple case of a
fluid in steady motion through a straight horizontal pipe of constant diameter,
illustrated by Fig. 2.9.
Flow is fully-developed, meaning that conditions do not vary between positions
along the pipe. Therefore, the shear stress exerted by the pipe walls on the fluid (τo) is
the same at all sections. Between the two sections A and B, a distance (L ) apart, the
total area of the pipe wall is πDL, so that the total force exerted by the pipe walls on
the fluid is πDLτ0. The linear momentum equation applied to this case of steady
uniform fully developed flow states that the total force on the fluid between sections
A and B must be zero (because the momentum flux across section A is equal to that
across section B), giving:
2.4 Basic Relations for Fluid Flow in Pipelines 25
πD2
ðpA- pB Þ - πDLτ0 = 0 ð2:13Þ
4
where pA and pB are the (static) pressures in the fluid at the two sections.
Rearranging Eq. 2.13:
dp pA - pB 4τ0
- = = ð2:14Þ
dL L D
Or:
ðpA - pB ÞD
τ0 = ð2:15Þ
4L
Equations (2.14) and (2.15) apply whether or not the fluid is Newtonian. Another
basic equation for angular momentum, needed for the analysis of centrifugal pumps,
will be presented in that context in Chap. 8.
The mechanical energy balance for a flowing, incompressible fluid is usually written
in the form known as Bernoulli’s equation, given below in terms of the head of the
fluid. Head is a concept used extensively by civil and mining engineers and
represents a measure of the mechanical energy of a flowing fluid per unit mass or
weight. It indicates the height by which the fluid would rise if the energy were
converted to potential energy and is expressed in units of length. In the past, this was
sometimes called the “total dynamic head,” especially in reference to the head
produced by a pump; however, this term is no longer widely used. The head of a
fluid of density ρf flowing at velocity V in a pipe at elevation z above a reference level
and at pressure p is given by
V2 p
H= þ þz ð2:16Þ
2g ρf g
In this equation:
• The first term (V2/2g) represents the kinetic energy of the fluid, often called the
velocity head.
• The second term (p/ρfg) represents the static pressure energy in the fluid, or the
pressure head.
• The final term (z) represents the elevation head, often referred to as the static
head.
26 2 Review of Fluid and Particle Mechanics
V 2B - V 2A pB - pA
H PUMP = H B - H A = þ þ ðzB-zA Þ ð2:17Þ
2g ρf g
In terms of the total volumetric flow rate (Q) and the pipe diameters at the two
sections (DA and DB) (Eqs. (2.9) and (2.10)), this becomes
8Q2 1 1 p - pA
H PUMP = 2 - 4 þ B þ ðzB- zA Þ ð2:18Þ
π g DB DA
4 ρf g
The second term on the right side of Eq. (2.18), the pressure head, is normally by
far the largest. The static head term (zB - zA) is usually relatively small, and it
depends on the pump geometry, orientation, and dimensions. Centrifugal pumps are
sometimes made with different suction and discharge diameters (see Chap. 8) so that
the first or velocity head term can become significant at large flow rates.
Now consider a section of pipeline. While a pump increases the head of the fluid
passing through it, the fluid flowing in a pipe experiences a decrease in head due to
energy losses associated with pipe wall friction and turbulent eddies at obstructions
and discontinuities in the flow, such as valves or elbows. The reduction in head per
unit length of pipe is known as the hydraulic gradient. It is expressed as height of
head lost per length of pipe run, for example (meters head/meters pipe), so its
numerical value is independent of the system of units used. For a fluid flowing in
a straight horizontal pipe of constant diameter (as in Fig. 2.9), the only head term that
2.4 Basic Relations for Fluid Flow in Pipelines 27
varies along the pipe is that arising from changes in pressure and the hydraulic
gradient is expressed as
pA - pB
jf = ð2:19Þ
ρf gL
where ρf is the density of the fluid flowing in the pipe. Note that in slurry applica-
tions, this fluid is a mixture of liquids and solids, and the density used in Eq. 2.19 is
the mixture density. This will be denoted as ρm in subsequent chapters to distinguish
it from the density of the carrier fluid. The associated slurry hydraulic gradient will,
in a similar way, be denoted by jm.
Many of the models used for determining slurry hydraulic gradients are actually
determining a pressure gradient, but for theoretical convenience and easy conversion
to a true hydraulic gradient, this pressure gradient is expressed in the form of a
hydraulic gradient, but with the density of the slurry replaced by the density of
standard water (ρw) = 1000 kg/m3. This modified hydraulic gradient, which may
also be called a dimensionless pressure gradient, is expressed as height of standard
water per length of pipe, for example (meters standard water/meter pipe), and is
denoted by “i” instead of “j”:
pA - pB
if = ð2:20Þ
ρw gL
As with j, values of i associated with the slurry mixture will often be denoted as
im. It is important to emphasize that i is proportional to the pressure gradient,
regardless of the density of the slurry being pumped, while j is proportional to the
head gradient, which incorporates the density of the slurry mixture. Aside from its
uses in the formulation of theoretical models, i is useful for comparing the energy
required to overcome pipeline friction losses at various operating conditions. On the
other hand, j is useful when comparing the form of friction loss curves for various
operating conditions, and also when examining the interaction between the pump
and the pipeline characteristics, since the head produced by a centrifugal pump
remains relatively constant regardless of the density of the slurry mixture, with
some corrections as discussed in Chap. 9.
A simple illustration of the significance of jf is given in Fig. 2.11.
If we imagine that “sight glasses” are attached to the pipe (i.e., vertical, open-
ended, transparent tubes), then the height to which the fluid rises in each sight glass
shows the pressure head inside the pipe at that point. For steady, fully developed
pipe flow, with constant hydraulic and pressure gradients, the levels in the sight
glasses set along the pipe will lie on a straight line. This is known as the hydraulic
grade line, and its inclination, or slope, is the hydraulic gradient. If the fluid in the
sight glasses is the same as the fluid flowing in the pipe, then the hydraulic gradient is
jf. If the fluid in the sight glasses is standard water and we assume, for the sake of this
thought experiment, that the two fluids do not mix, then the hydraulic gradient is if.
28 2 Review of Fluid and Particle Mechanics
1.0
jf
A B C
pA
A
The two gradients differ for the same flow due to different heights caused by the
different densities of the fluids in the sight glasses.
Consider now a pipe at an angle to the horizontal, as shown in Fig. 2.12.
The pressure difference is measured between sections A and B by connecting the
pipe at these points to a manometer or a differential pressure sensor. Point B is Δz
above A. The sensor is h above A, and therefore (Δz-h) below B, and the connections
from the sensor to the pipe are filled with the same fluid as in the pipe, of density ρf.
If the pressure in the pipe at A is pA, then the pressure on the upstream side of the
sensor is lower by the hydrostatic column h:
p0A = pA - ρf gh ð2:21Þ
In other words, the sensor will measure the change in the combination of pressure
head and static head between A and B, which gives the pressure losses due to friction
alone. (The decrease in pressure due to elevation is not included in the pressure
difference measurement.) For the case where the pipe is of constant section, this
measurement is equal to the change in total head, i.e. jfρfgL where L is the distance
from A to B. Thus, a simple manometric measurement, such as that illustrated in
Fig. 2.12, serves to measure the hydraulic gradient.
Using Eq. (2.14), the hydraulic gradient is related to the wall shear stress (τo):
4τ0
jf = ð2:24Þ
ρf gD
Standard methods for predicting the wall shear stress and hydraulic gradient for
Newtonian fluids are given in the following section.
Strictly speaking, the hydraulic gradient, as a measure of the energy head loss,
represents the slope of the energy grade line. That line goes parallel to the hydraulic
grade line if the pipe has a constant diameter; hence, the velocity head does not vary
along the length of the pipe.
Note that the definitions of if and jf incorporate the local acceleration due to
gravity. If the flow takes place on another planet or in a system subjected to other
acceleration, their values would be different even if τo and the pressure gradient were
unchanged.
In the previous section, Fig. 2.9 shows a fluid flowing along a straight horizontal
cylindrical pipe under the action of the pressure gradient dp/dL, where L is distance
measured along the pipe. From Eq. (2.15), the shear stress at the pipe wall is
D dp
τ0 = - ð2:25Þ
4 dL
A similar force balance on the fluid within the cylinder of radius (r), coaxial with
the pipe, shows that the shear stress varies across the pipe section, as shown in
Fig. 2.13, and at radius r, the shear stress is
r dp
τ= - ð2:26Þ
2 dL
30 2 Review of Fluid and Particle Mechanics
t0
t laminar
u
r
turbulent
To obtain the velocity profile, Eq. (2.28) is integrated with the two conditions:
• Axial symmetry, meaning the velocity gradient is zero on the pipe axis (i.e.,
du/dr = 0 at r = 0).
• The no-slip condition, meaning the velocity is zero at the wall (i.e., u = 0 at
r = D/2).
The resulting velocity profile takes the characteristic parabolic form, shown in
Fig. 2.13 and given by
2.5 Pipeline Friction of Newtonian Fluids 31
4r 2
u = umax 1- 2 ð2:29Þ
D
In Eq. (2.29), umax is the maximum velocity in the pipe, which occurs on the axis
and is given by
Dτ0 D2 dp
umax = = - ð2:30Þ
4μ 16μ dL
Using Eq. (2.12), the total flow rate in the pipe is obtained as
πD2
Q= u ð2:31Þ
8 max
4Q u
V= = max ð2:32Þ
πD2 2
And so, for a Newtonian fluid in laminar flow in a cylindrical pipe, the maximum
velocity is twice the mean velocity. Finally, from Eq. (2.30) and (2.31):
πD4 dp
Q= - ð2:33Þ
128μ dL
and
8V
τ0 = μ ð2:34Þ
D
turbulent core
y+ = 70
buffer layer y+ = 11.6
linear sub-layer y+ = 5
wall and κ is von Karman’s coefficient. The value κ = 0.4 is often employed and will
be used here, giving the velocity gradient as 2.5 U*/y.
The local velocity (u), obtained by integration, varies with the logarithm of y. If
the pipe wall is hydraulically smooth, the velocity distribution is given by
u yU
= 2:5 ln þ 5:5 ð2:35Þ
U ν
where ln indicates the natural logarithm and ν is the kinematic viscosity = (μ/ρf). The
left-hand side of this equation is a dimensionless velocity, denoted u+, while the ratio
yU*/v is a dimensionless distance from the wall, denoted y+.
At small values of y+, the logarithmic velocity law of Eq. (2.35) will not apply,
since immediately adjacent to a smooth wall, there is a viscous sublayer where
viscous stresses dominate over Reynolds stresses, and the flow conditions can be
considered laminar in nature.4 Therefore, by Eq. (2.1), du/dy must equal τo/μ, which
results in a linear variation of u with y, equivalent to the statement that u+ equals y+ at
small values of y+. The viscous sublayer extends from the smooth wall at y+ = 0 out
to about y+ = 5, where turbulence first begins to be felt (see Fig. 2.14). The thickness
δ of the viscous sublayer is given by:
ν 5μ
δ=5 = ð2:36Þ
U ρf U
4
In reality coherent structures periodically penetrate this region forming turbulent wall streaks,
particle ejections and other phenomena (see e.g. Rashidi et al. (1990), Hetsroni (1993)) but these are
insufficient to change the overall linear velocity profile in this region making the region appear to be
laminar.
2.5 Pipeline Friction of Newtonian Fluids 33
is sufficient to use a simplified treatment in which the buffer layer is neglected, and
transition to turbulent behavior is assumed to occur abruptly at the point where
Eq. (2.35) intersects the linear velocity relation applicable near the wall. This occurs
where both y+ and u+ equal 11.6, so the linear viscous sublayer is assumed to extend
from the wall to y+ = 11.6, and Eq. (2.35) is used for all larger values of y+.
As the axis of the pipe is approached, the observed velocity profile diverges
somewhat from the logarithmic law of Eq. (2.35), and detailed models of turbulent
flow take this divergence into account. For present purposes, the logarithmic law is
adequate for the entire flow outside of the viscous sublayer. Since this sublayer
usually occupies a very small portion of the pipe area, a very close approximation of
the average velocity (V ), which is the total flow rate divided by the pipe section area
πD2/4, can be obtained by integrating Eq. (2.35). The result may be written as
follows:
V U D
= 2:5 ln ð2:37Þ
U ν
Both dimensionless groups in this equation merit careful consideration. The ratio
V/U* can be expressed directly in terms of the dimensionless group known as the
friction factor. This book uses the Darcy-Weisbach (Moody) form of the friction
factor, defined as follows:
8τ0
f= ð2:38Þ
ρf V 2
This is the definition commonly used by civil and mechanical engineers, whereas
chemical engineers may be more familiar with the Fanning friction factor, equal to
f/4. From Eq. (2.38), it follows that V/U* = (8/f )1/2, or
rffiffiffi
f
U =V ð2:39Þ
8
The dimensionless quantity U*D/v found in Eq. (2.37) is called the shear Reyn-
olds number, denoted as Re*. Like the better-known pipe Reynolds number
(Re = VD/v or ρVD/μ), it gives an indication of the state of flow. For cases where
the pressure gradient, pipe diameter, and fluid properties are known, U* and Re* can
be determined immediately, and the mean velocity (V ) is found by substituting these
quantities into Eq. (2.37).
In cases where V and Re are known and U* (and the pressure gradient) are
required, we can solve for these by re-ordering the above relationships:
34 2 Review of Fluid and Particle Mechanics
y+ = 70 y+ = 70
buffer layer buffer layer
y+ = 5 y+ = 5
linear sub-layer linear sub-layer
(a) (b)
turbulent core
y+ = 70
buffer layer
y+ = 5
linear sub-layer
(c)
Fig. 2.15 Interaction of asperities with flow in boundary layers: (a) hydraulically smooth condi-
tion: ε+ < 5; (b) transition: 5 ≤ ε+ ≤ 70; (c) fully rough condition: ε+ > 70
rffiffiffi rffiffiffi
UD f VD f
Re = = = Re ð2:40Þ
ν 8 ν 8
With Re known, this equation can be solved for f. Although iteration is required,
the range of f is small, and the solution can be obtained quickly.
Note that Eq. (2.41), like Eq. (2.37), from which it is derived, applies only to
turbulent flow with hydraulically smooth pipe walls. A smooth wall does not require
asperities to be completely absent, only that the size of the typical asperity or
roughness (ε) is too small to penetrate the viscous sublayer and influence the
turbulent portion of the flow. For larger values of ε, the relative roughness (ε/D) is
a significant parameter influencing pipe friction. For fully rough pipes, the viscous
sublayer and buffer layer are completely hidden between the asperities on the pipe
wall, so that the roughness interacts directly with the turbulent flow. In this case,
viscosity is no longer important, and the friction equation depends only on ln(D/ε),
instead of ln(Re*). Figure 2.15 shows the various cases from smooth to fully rough in
terms of the dimensionless size of asperity ε+ = εU*/ν.
The Colebrook–White equation provides an appropriate transition function that
incorporates both “smooth” and “rough” behavior as limiting cases. Rearranged
somewhat (Streeter and Wylie 1975), the equation may be expressed as follows:
2.5 Pipeline Friction of Newtonian Fluids 35
rffiffiffi
V 8 ε=D 2:51
= = - 2:43 ln þ pffiffiffi ð2:42Þ
U f 3:7 Re f
Here, the inverse of the von Karman coefficient is assigned a value of 2.43, which
is not significantly different from the 2.5 value used in other friction equations.
As was the case for Eq. (2.37), if the pressure gradient is known (together with D, ε,
and v), Eq. (2.42) allows the mean velocity to be calculated directly. If Re is
known, together with ε/D, the equation can readily be iterated for f.
Many explicit approximation formulae for f have been developed to avoid the
iterations associated with Eqs. (2.41) and (2.42). The Swamee-Jain (Swamee and
Jain 1976) formula solves directly for f in any turbulent flow:
0:25
f=h i2 ð2:43Þ
log ε=D
3:7 þ 5:74
Re 0:9
The Churchill formula (Churchill 1977) gives a direct solution for f in laminar,
transition, and turbulent flows:
" #121
12
8 1
f =8 þ ð2:44Þ
Re ðA þ BÞ1:5
where
" !#16 16
1 37530
A = 2:457 ln ε=D and B = :
Re
3:7 þ Re 0:9
5:76
Both of these explicit formulations agree with Eq. (2.42) to within a percentage
or so.
The same result can be obtained directly from a graph of the relationship given by
Eq. (2.43). This plot is often known as the Moody diagram or the Stanton-Moody
diagram, as shown in Fig. 2.16. Figure 2.17 displays the portion of the diagram that
is of most interest in industrial pipeline applications.
Various regions can be distinguished on the Moody diagram. For Re < 2000, f is
independent of roughness and is given by
64
f= ð2:45Þ
Re
This range corresponds to laminar flow, and Eq. (2.45) is simply Eq. (2.34)
written in terms of the dimensionless groups introduced above.
For Re between about 2000 and 3000, flow can be laminar or turbulent. In
industrial practice, it will almost always be turbulent. For Re > 3000, flow is
36 2 Review of Fluid and Particle Mechanics
0.020
0.001
0.0008
0.018
0.0006
0.016 0.0004
e /D (-)
f (-)
0.014
0.0002
0.012 0.0001
0.00005
smooth
0.010
105 106 0.000005 107
Reynolds number (-)
Fig. 2.17 Pipe friction factor in normal operating range. Red symbol is a result of Example 2.1
2.5 Pipeline Friction of Newtonian Fluids 37
turbulent, and f depends on both Re and ε/D. The curves in Fig. 2.17 are each drawn
for one value of the relative roughness. In general, the friction factor decreases as Re
increases and as roughness decreases. However, for sufficiently large values of Re,
the buffer layer becomes thinner than the asperities so that the range indicated as
fully rough is entered: the horizontal curves show that f now depends on ε/D but is
independent of Re. Even in the transitional rough range, the dependence of f on Re is
weak. Thus, for turbulent flow in a given pipe with fixed relative roughness, it is
frequently sufficient to treat the friction factor as a constant, characteristic of
the pipe.
Detailed characterization of roughness is a subject in itself, as the microscopic
geometry will differ from one rough surface to another. However, commercial pipes
are commonly characterized in terms of the equivalent sand grain roughness,
wherein their friction characteristics are compared with systematic measurements
obtained by gluing sand grains to the walls of test sections of pipes (Nikuradse
1933). As an example, the equivalent sand grain roughness for commercial steel pipe
is about 45 μm. It must be remembered that the effective roughness can change in
service: corrosion can increase ε, while the polishing action of the particles in a
slurry can reduce ε (and possibly also increase D if erosion is severe). For the flow of
slurry in a steel pipe, the roughness of smooth worn steel (1.5 μm) is typically used.
For a given flow of a given fluid in a given pipe, it is possible to calculate Re and
ε/D, and hence obtain f from Fig. 2.17 or Eq. (2.43). From Eq. (2.19) and Eq. (2.38),
the hydraulic friction gradient follows as:
V2 ρf V 2
jf = f and if = f ð2:46Þ
2gD ρw 2gD
As f is approximately constant for turbulent flow in a given pipe, Eq. 2.46 shows
that the hydraulic gradient varies roughly as V2 (or as Q2).
The approach to estimating the hydraulic and pressure gradients jf and if
(as summarized above and used throughout this book) gives results that are essen-
tially equivalent to those of other, more obviously empirical methods. An example of
such methods is the “C-factor” of Hazen and Williams, which is still sometimes used
in the mining industry. However, the use of f, with its dependence on Reynolds
number and relative roughness, is preferred because it gives an indication of flow
conditions and is more readily extended to slurry flow interpretation.
The loss of head associated with fittings, such as elbows, and valves is usually
estimated by multiplying the velocity head by a loss coefficient. For turbulent flows,
the loss coefficient does not change with the Reynolds number, but for laminar
flows, it is higher than the turbulent flow coefficient and is dependent on the
Reynolds number (Herwig et al. 2010). The loss coefficient given in most texts is
valid only for turbulent flow. The most widely used loss coefficients are 0.5 for a
standard 90 degree elbow (0.2 for a long-radius elbow) and 0.8 for an abrupt
(unrounded) entry, although the coefficient varies with pipe diameter and bend
radius of curvature. At a pipe exit, the full velocity head is lost, and this loss is
38 2 Review of Fluid and Particle Mechanics
150m
2
Dz
1 45m
10m
B
A 300m
sump
P = ρf gjf QL ð2:47Þ
which can be rearranged, using Eq. (2.9) and Eq. (2.46), as follows:
8ρf Q3 f
P= L ð2:48Þ
π 2 D5
The very strong dependence on pipe diameter is worth noting. If power con-
sumption is a major consideration, the economic incentive is towards using pipes of
large diameter. In later chapters, we address the effect of this consideration on slurry
system design.
Also in later chapters, we will examine the operability of piping systems that use
centrifugal pumps by matching the system pipe characteristic with the pump char-
acteristic. Figures 2.18 and 2.19 illustrate this idea for a system conveying a liquid
alone. For a simple piping system shown schematically in Fig. 2.18, the total head
required varies with flow rate, as shown by curve 1 in Fig. 2.19. The static lift term,
Δz, from the surface in the sump to the pipe discharge, does not depend on flow rate.
The flow-dependent terms include the friction losses in the pipe and fittings and also
the velocity head, which changes as the velocity in the pipe changes (e.g., due to
diameter changes) and which is often lost when the piping system discharges into a
tank, pond, or piece of process equipment.
2.5 Pipeline Friction of Newtonian Fluids 39
head
B
A
2
Δz
flow rate
4Q 4 0:12
V= = = 3:72 m=s
πD 2
π 0:20272
This mean velocity is in the range of operational velocities typically used for
coarse slurries in an 8-inch pipe.
For water at 20 °C, ρf = 998.2 kg/m3 and μf = 1.002 × 10-3 Pa.s (Table 2.1).
Therefore, the pipe Reynolds number is
40 2 Review of Fluid and Particle Mechanics
ρf VD 998:2ð3:72Þ ð0:2027Þ
Re = = = 7:51 × 105
μ 1:002 × 10 - 3
0:25 0:25
f=h i2 = h i2 = 0:0153
ε=D
log 3:7 þ 5:74
Re 0:9
log 3:7 þ 751040
0:000222 5:74
0:9
The same result is obtained from Fig. 2.17. Values for the Darcy-Weisbach
friction factor in the range of 0.01 to 0.02 will prove to be fairly typical for water
alone pumped at conditions commonly used for coarse slurries.
(c) Hydraulic Gradient
From Eq. (2.46), the hydraulic gradient is
V2 0:0153ð3:72Þ2
jf = f =
2gD 2ð9:81Þð0:2027Þ
= 0:0532 ðmeters head of flowing fluid per meters of pipeÞ and
ρf 998:2
if = j = 0:053
ρw f 1000
= 0:0531 ðmeters head of standard water per meters of pipeÞ
Values of the order of a few meters head per hundred meters of pipe are again
typical of water pumped at conditions appropriate for a coarse slurry.
(d) Wall Shear Stress, Shear Velocity, and the Thickness of the Viscous Sublayer
From Eq. (2.38), the wall shear stress is
f 0:0153ð998:2Þð3:72Þ2
τ0 = ρf V 2 = = 26:4 N=m2
8 8
rffiffiffi rffiffiffiffiffiffiffiffiffiffiffiffiffiffi
f 0:0153
U =V = 3:72 = 0:163 m=s
8 8
The thickness of the viscous sublayer is estimated from Eq. (2.36) as follows:
μ 5ð1:002 × 10 - 3 Þ
δ=5 = = 3:1 × 10 - 5 m
ρf U 998:2 0:163
This value (i.e., 31 μm) can be compared with the equivalent sand grain rough-
ness of 45 μm. As expected for a flow that is well into the transitional rough range
(see Fig. 2.16), the sublayer thickness and the equivalent roughness are of compa-
rable magnitude, and the roughness asperities penetrate through the sublayer into the
buffer layer. The value for sublayer thickness, a few tens of microns, is typical and is
worth noting.
(e) Head Required from the Pump
To determine the head required for the pump in the simple pipe-pump system in
Fig. 2.18, apply the Bernoulli equation to both the suction and discharge sides of the
system.
• The suction side has the water surface in the reservoir as the inlet profile (point 1)
and the end of the suction pipe (i.e., the suction inlet of the pump) as the outlet
profile (point A). The water surface in the reservoir is 10 meters above the suction
pipe outlet.
• The inlet to the discharge pipe (point B) is considered to have the same elevation
as the suction pipe outlet. The outlet of the discharge pipe (point 2) is 45 m higher.
The suction pipe and the discharge pipe have the same internal diameter (D) and
pipe wall roughness. System fitting losses (minor losses) consider the loss
coefficient K = 0.2 for the long-radius elbows and K = 0.5 for the pipe entrance.
The gauge pressure at the pump suction inlet is determined from the Bernoulli
equation (Eq. 2.16) applied to the suction pipe:
pA - patm V2
= Δz1-A - - H loss,1-A
ρf g 2g
The gauge pressure at the pump discharge results from the Bernoulli equation for
the discharge pipe:
pB - patm
= ΔzB-2 þ H loss,B-2
ρf g
8Q2 1 1 p - pA p - pA
H PUMP = 2 - þ B þ ðzB- zA Þ = B
π g D4B D4A ρf g ρf g
Using the above equations for the static pressures pA and pB:
V2
H PUMP = Δz1-2 þ þ H loss,1- 2 ,
2g
in which:
X V2
L
H loss,1 - 2 = H frict,1 - 2 þ H minor,1 - 2 = f 1-2 þ K
D 2g
and thus:
The properties of slurries depend very strongly on the tendency of the particles to
settle out from the conveying liquid. For transport of settling slurries (for their
definition, see Chap. 3), an important parameter is terminal velocity (vt), the velocity
at which a single particle settles through a large volume of quiescent liquid. The
terminal velocity depends on the liquid properties (ρf and μ) and on the particle
diameter (d ), its density (ρs), and, to a lesser extent, its shape.
For vertical flow of settling slurries, the hindered settling velocity is also impor-
tant. This is the reduced settling velocity experienced by collections of particles that
are able to interact physically or hydraulically with each other as they settle.
When particles have fully settled, their concentration, as achieved without
compacting or vibrating the sediment, is referred to in later chapters as the loosely
packed volume fraction and is denoted Cvb.
Particle sizes are commonly reported as “screen size” (referring to the opening in
a standard sieve or screen). Standard screen series have often been expressed as
“mesh size” based on the number of openings per inch, but openings in mm (or μm)
are now in common use. Particle size distributions are reported as the fraction
(by mass) of particles passing through each screen in a series of successively smaller
openings.
In general, although particles in slurries are not spherical, the sphere represents a
convenient reference case in the analysis. Figure 2.20 shows the forces acting on a
rigid sphere settling through a fluid.
In this figure, the weight of the particle is partially reduced by the buoyancy of the
surrounding fluid. When the spherical particle is moving steadily at its terminal
44 2 Review of Fluid and Particle Mechanics
vts
−
6
velocity (vts), the resulting immersed weight (also called submerged weight or net
weight) is balanced by the drag of the fluid:
πd3
FD = ðρs - ρf Þg ð2:49Þ
6
where vts refers specifically to the terminal velocity of a spherical particle. Given that
there are five parameters and the usual three basic dimensions of mass, length, and
time, Buckingham’s theorem shows that two dimensionless groups suffice to express
Eq. (2.50) in a general form. Most commonly, the two dimensionless groups selected
are
8F D
drag coefficient : C D = ð2:51Þ
πd2 v2ts ρf
ρf vts d
particle Reynolds number : Re p = ð2:52Þ
μ
The function indicated in this equation has been fitted to the many experimental
determinations of drag or terminal velocity, to give the “standard drag curve” shown
in Fig. 2.21.
2.6 Settling of Solids in Newtonian Fluids 45
102
10
CD (-)
0.1
0.01
1
1.E+00 10
1.E+01 1.E+022
10 1.E+033
10 1.E+044
10 1.E+055
10 1.E+066
10 1.E+07 7
10
Rep (-)
For Rep < 2.6 × 105, which amply covers the range encountered in most slurries,
the curve is approximated well by an expression given by Turton and Levenspiel
(1986):
24 0:413
CD = 1 þ 0:173 Re p 0:657 þ ð2:54Þ
Re p 1 þ 1:63 104 Re p - 1:09
The terminal velocity can then be determined from Eq. (2.50) and Eq. (2.54), but
the procedure is iterative, with successive estimates for vts updated to converge on
the solution.
The reasons for the form of the curve in Fig. 2.21 are discussed in detail by Clift
et al. (1978). For low particle Reynolds numbers (such as, for example, Rep < 0.1),
Stokes’ law applies, with the drag force given by the theoretical result:
F D = 3πμvts d ð2:55Þ
d2 ðρs - ρf Þg
vts = ð2:56Þ
18μ
At much larger Reynolds numbers, in the approximate range 750 < Rep < 3 × 105,
the drag coefficient is roughly constant and close to 0.445. This range is known as
the “Newton’s law” range, based on Newton’s experiments with falling objects (e.g.,
46 2 Review of Fluid and Particle Mechanics
inflated pigs’ bladders falling within the dome of St. Paul’s Cathedral, London; see
Newton (1726)). In this range, the terminal velocity of a sphere falling through water
can be calculated using:
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ρ - ρf
vts = 1:73 gd s ð2:57Þ
ρf
As a general guide for sand-density particles in water, Stokes’ law applies for
particles smaller than about 50 μm, while Newton’s law applies for particles larger
than about 2 mm.
As noted previously, two independent dimensionless groups are needed to
express empirical drag results in general form. These groups can be selected for
convenience, however. Various possibilities and their uses are reviewed by Clift
et al. (1978). Following the suggestion of Grace (1986), we find it convenient to
define a dimensionless particle diameter d* (equal to the cube root of what is often
called the Archimedes number):
!1=3 1=
3C D Re2p ρf ðρs - ρf Þg 3
d = =d ð2:58Þ
4 μ2
1= 1=3
Rep 4Rep 3 ρ2f
vts = = = vts ð2:59Þ
d 3C D μðρs - ρf Þg
and calculation of vts* (hence, vts), requires no iteration. This functional relation has
been worked out in considerable detail and is entirely suitable for particles falling in
Newtonian fluids. For example, Haider and Levenspiel (1989) give the following for
spherical particles:
"
0:824 0:412 i - 1:214
18 0:321
vts = þ ð2:61Þ
d 2 d
However, for the analogous case in non-Newtonian fluids (see Sect. 2.7), diffi-
culties arise because the viscosity is included in both dimensionless variables. To
cover both Newtonian and non-Newtonian cases, an alternative method has been
developed, as described by Wilson et al. (2003) and Horsley et al. (2003). This
method is based on a pair of dimensionless variables that employ concepts devel-
oped in the pipe-flow analysis of Prandtl (1933) and Colebrook (1939).
2.6 Settling of Solids in Newtonian Fluids 47
The method expresses the velocity ratio (mean velocity to shear velocity) as a
function of the shear Reynolds number (basedpon shear
ffi velocity rather than mean
ffiffiffiffiffiffiffiffiffiffi
velocity). In pipe flow the shear velocity, U = τ0 =ρf. (Note that τ0 is uniform for a
circular pipe.) The shear stresses that develop over the surface of a spherical particle
are non-uniform, but the mean surficial shear stress, denoted, forms a useful basis for
analysis. This stress is defined as the submerged weight force divided by the surface
area of the sphere, which is πd2, where d is the sphere diameter. The submerged
weight force is the product of the sphere volume πd3/6 and (ρs - ρf)g, where g is the
gravitational acceleration and ρs and ρf are the densities of the solid and fluid phases,
respectively. Thus, the mean surficial shear stress is given by:
ðρs - ρf Þgd
τ = ð2:62Þ
6
For the falling particle case, the shear velocity is based on. Thus:
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ðρs - ρf Þgd
V = ð2:63Þ
6ρf
which can also be expressed in terms of relative density or specific gravity (S), of the
solids (Ss = ρs/ρw) and carrier fluid (Sf = ρf/ρw), where ρw is the density of standard
water:
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ðSs - Sf Þgd
V = ð2:64Þ
6Sf
For pipe flow, the velocity ratio is the mean velocity divided by V*. For falling
particles, the analogous ratio is based on the terminal fall velocity of the particle (vt),
giving vt/V* as the velocity ratio. The particle shear Reynolds number (Re)* has
the form:
ρf V d
Re = ð2:65Þ
μ
where d is the particle diameter (analogous to the pipe diameter for the pipe-flow
case) and, for a Newtonian fluid, μ is the dynamic viscosity.
Next, we consider the form of the settling curve for Newtonian fluids for the vts/V*
and Re* axes. In terms of these variables, the drag coefficient equals 8/[vts/V*]2 and
the conventional Reynolds number Re equals the product of Re* and vts/V*. For Re
larger than about 1100, the drag coefficient can be taken as effectively constant at
0.445, equivalent to vts/V* = 4.24 for Re* > 260. At the low end of the Reynolds
number range (Re* < 1), settling obeys Stokes’ law which can be expressed as vts/
V* = Re*/3. In the intermediate region, the coordinates of individual points on the
48 2 Review of Fluid and Particle Mechanics
10
0.1
0.01
0 .1 1 10 100 1,000 10,000
Re* (-)
Fig. 2.22 Curve of relative fall velocity versus shear Reynolds number. (From Wilson et al. 2003)
CD - Re curve can be transformed into vts/V* and Re* values. When these values are
plotted on the new curve, the result is the shape shown in Fig. 2.22.
This complete curve may be approximated as:
- 0:9362
vts 3:5543
= 0:2181 þ ð2:66Þ
V Re
h i-1
18 2:3348 - 1:7439ψ
vt = þ ð2:67Þ
d2 d0:5
rffiffiffiffiffiffiffiffi
4Ap
da = ð2:68Þ
π
For particles whose sizes are determined by sieving rather than by microscopic
analysis, da is slightly smaller than the mesh size. However, unless the particles are
needle-shaped, the difference between da and the screen opening is relatively small,
generally less than 20%.
The shape of the particle is described by the volumetric shape factor (K )
defined as:
volume of particle
K= ð2:69Þ
d3a
so that K is 0.524 for a sphere. Table 2.7 lists representative values for various
mineral particles. In general, the more angular or flaky the particle, the lower the
value of K.
The procedure for calculating the terminal velocity begins by using the method
presented above to calculate vts for the sphere of diameter da and the same density as
the particle of interest. The value for the non-spherical particle is then given by
vt = ξvts ð2:70Þ
where the velocity ratio (ξ) is a function of the volumetric shape factor (K ) and a
weak function of the dimensionless diameter (d*). Values for ξ are obtained from the
curves in Fig. 2.23.5 It will be seen that, for common particles like sand and coal, the
5
The curves in this figure can be approximated by the function
ξ = 0:1268 þ 1:4K þ 0:09 exp - 0:5ðlnðd =14Þ=0:7Þ2
50 2 Review of Fluid and Particle Mechanics
0.2
0.1
0.1
1 10 100
d* (-)
Fig. 2.23 Ratio of terminal velocity of non-spherical particle to the value for a sphere (ξ) as a
function of dimensionless diameter (d*)
terminal velocity is typically 50–60% of the value for the equivalent sphere. The
value of ξ allows for the lower volume (and therefore the lower immersed weight) of
the particle compared to the sphere, and also for differences in drag.
The hindered settling velocity (vt´) is normally less than vt and is strongly
dependent on the volume concentration of solids. It can be estimated with reasonable
reliability by the correlation of Richardson and Zaki (1954):
v0t = vt ð1 - Cv Þn ð2:71Þ
Equation (2.71) allows for two phenomena that reduce vt´ below vt: the increased
drag caused by the proximity of other particles; and the up-flow of liquid as it is
displaced by the descending particles. The index (n) depends on d*, and it is larger
for particles settling in the range of Stokes’ law (n = 4.6) and smaller in the range of
Newton’s law (n = 2.4). For intermediary values in the range 0.01 < Rep < 1000, the
following formulation can be used (Rowe 1987):
2:35 2 þ 0:175 Re 0:75
p
n= ð2:72Þ
1 þ 0:175 Re 0:75
p
2.6 Settling of Solids in Newtonian Fluids 51
Estimate the terminal velocity and hindered settling velocities of sand particles with
a shape factor of 0.26 in water at room temperature. Consider particles of sizes 0.2,
0.5, 1.0, and 2.0 mm. (Chap. 6 will show that particle settling is of the greatest
significance for heterogeneous flow, which applies to particles in this size range.)
(a) Terminal Velocities of Equivalent Spheres
For water at 20 °C, ρf = 998.2 kg/m3 and μ = 1.002 × 10-3 Pa.s (from Table 2.1).
The density of quartz sand (ρs) is typically 2650 kg/m3. Therefore, the shear velocity
based onqthe meanffi surficial shear stress (Eq. (2.63)) is:
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffi
V = ðρs -6ρρf Þgd or 2:71d with d in meters. Table 2.8 shows the calculated
f
values that were obtained for this quantity for the various particle sizes. The table
also shows the values of Re* (i.e., ρfV*d/μf). This quantity is used to calculate vts/V*
using Eq. (2.66). Multiplying the ratio vts/V* by V* gives the fall velocity for a
spherical particle (vts), which is shown in Table 2.8. Note that the 0.2 mm particles
lie well beyond Stokes’ law range, while the 2 mm particles are not quite into
Newton’s law range.
(b) Terminal Velocities of Sand Grains
To obtain the terminal fall velocity of the sand particles, first multiply the values
of vts by the velocity ratio (ξ), which, in turn, depends on the dimensionless particle
size (d*), given by Eq. (2.58):
" #1=3
h i1=3 998:2ð2650 - 998:2Þ 9:81
d = ρf ðρs - ρf Þg d or d = 2 d = 2:526 104 d
μ2
1:002 10 - 3
52 2 Review of Fluid and Particle Mechanics
The values of the velocity ratio (ξ) are obtained from Fig. 2.23 by interpolation
between the contours for K = 0.2 and 0.3 to K = 0.26, at the appropriate value of d*
in each case. Using the values for ξ and vts from part (a) gives the results shown in
Table 2.8. Note that the variation of ξ over the range of particle sizes is rather small.
(c) Hindered Settling Velocities of Sand Grains
The hindered settling velocity of sand particle collections of the volume concen-
tration Cv = 0.2 is determined using the Richardson-Zaki formula (Eq. 2.71). It
requires the index n (Eq. 2.72), which depends on the particle Reynolds number
Rep = ρfvtd/μ. The results are included in Table 2.8.
While several books and correlations have been published on particle settling in non-
Newtonian fluids (Chhabra 1992; Michaelides 2006), these describe the motion of
particles in stationary fluids. This is of little use in the pumping of slurries and
suspensions where the fluids are sheared, and where the fluid viscosity changes due
to the spatial variation in fluid velocity profiles (i.e., shear rates).
Consider the case of a particle immersed in a stationary fluid that has a yield stress
(Fig. 2.24).
For the particle to settle, the fluid surrounding it must yield. By performing the
same force balance as before, but also including the surficial force obtained through
integration of the yield stress over the particles surface, we can show that if the
fluid’s yield stress exceeds a critical value (τyc), the particle will not settle (i.e.,
vts = 0). This critical yield stress is given by:
where k = 3/(2π) for spherical particles and ~ 0.1 for normal crushed rock.
−
6
2.7 Settling of Solids in Non-Newtonian Fluids 53
The particle will settle if the yield stress is less than this limiting value, but it will
be at a slower rate than if the particle were immersed in an equivalent Newtonian
fluid. This is due to subtle changes in the flow field around the particle and the
addition of unsheared regions of fluid attached to the poles of the particle (Ansley
and Smith 1967; Atapattu et al. 1995; Beris et al. 1985).
When the fluid is sheared, however, all motion around the particle which alters
the local fluid viscosity must be accounted for. This includes the local fluid velocity
profile, any spin that the fluid is undergoing, and the motion of the particle itself, as
shown in Fig. 2.25.
By summing these individual shear rates, we can deduce the shear rate of the fluid
local to the particle and hence its viscosity. Figure 2.3b shows that fluids with yield
stress are extremely shear-thinning at low shear rates. Thus, even for the smallest
shear rates, the local viscosity is substantially reduced, and the particle will settle,
albeit slowly. Tests conducted in recent years (e.g., Pullum 2017; Talmon and
Huisman 2005; Wilson and Horsley 2004) have shown that when this local viscosity
is used, particles settling in non-Newtonian fluids appear to obey the same, or
similar, drag laws as those in Newtonian fluids.
Figure 2.26 shows that particle settling in slurries, typical of those found in
thickened tailings disposal systems, obey Stokes’ drag law and that given by
Eq. (2.54), providing the local viscosity i.e. that of the fluid next to the particle,
is used.
54 2 Review of Fluid and Particle Mechanics
Fig. 2.26 Reynolds number drag law relationship for solids settling in non-Newtonian fluids.
(Matoušek et al. 2021)
References
T. Allen, Particle size measurement, 2nd edn. (Chapman and Hall, London, 1975)
R.W. Ansley, T.N. Smith, Motion of spherical particles in a Bingham plastic. AICHE J. 13,
1193–1196 (1967)
B.C. Aschenbrenner, A new method of expressing particle sphericity. J. Sediment. Petrol. 26,
15–31 (1956)
D.D. Atapattu, R.P. Chhabra, P.H.T. Uhlherr, Creeping sphere motion in Herschel—Bulkley fluids:
Flow field and drag. J. Non-Newtonian Fluid Mech. 59, 245–265 (1995)
L.G. Austin, I. Shah, A method for inter-conversion of microtrac and sieve size distributions.
Powder Technol. 35, 271–278 (1983)
G.H. Bagheri, C. Bonadonna, I. Manzella, P. Vonlanthen, On the characterization of size and shape
of irregular particles. Powder Technol. 270(Part A), 141–153 (2015)
H.A. Barnes, A Handbook of Elementary Rheology (University of Wales Institute of
non-Newtonian Fluid Mechanics, Aberystwyth, 2000)
A.N. Beris, J.A. Tsamopoulo, R.C. Armstrong, Creeping motion of a sphere through a Bingham
plastic. J. Fluid Mech. 158, 219–244 (1985)
R.P. Chhabra, Bubbles Drops and Particles in Non-Newtonian Fluids (CRC Press Ltd., Boca
Raton, 1992)
R.P. Chhabra, J.F. Richardson, Non-Newtonian Flow in the Process Industries (Butterworth-
Heinemann, Oxford, 1999)
S.W. Churchill, Friction-factor equation spans all fluid-flow regimes. Chem. Eng. J. Nov. 97, 93–94
(1977)
R. Clift, J.R. Grace, M.E. Weber, Bubbles drops and particles (Academic Press Inc., London, 1978)
C.F. Colebrook, Turbulent flow in pipes, with particular reference to the transition region between
the smooth and rough pipe laws. J. Inst. Civ. Eng. 11 (1939)
R. Cooke, Thickened and paste tailings pipeline systems – Design procedure, part 2. Presented at
the 10th Intl. Seminar on Paste and Thickened Tailings, Perth, Australia, 2007 (2007)
References 55
J.R. Grace, Contacting modes and behavior classification of gas-solid and other two-phase suspen-
sions. Can. J. Chem. Eng. 64, 353–363 (1986)
A. Haider, O. Levenspiel, Drag coefficient and terminal velocity of spherical and non-spherical
particles. Powder Technol. 58, 63–70 (1989)
G. Hetsroni, The effect of particles on the turbulence in a boundary layer, in Particulate Two-Phase
Flow, ed. by M.C. Roco, (Butterworth-Heineman, Stoneham, 1993)
Horsley MR, Horsley RR, Wilson KC (2003) Non-Newtonian Effects on Fall Velocities of Pairs of
Vertically-Aligned Spheres. Presented at the Intl. Conf. on Non-Newtonian Rheometry, Uni-
versity of Wales, Cardiff, UK
I.J. Karassik, J.P. Messina, P. Cooper, C.C. Heald, Pump handbook, 4th edn. (McGraw- Hill,
New York, 2008)
L. Korson, W. Drost-Hansen, F.J. Millero, Viscosity of water at various temperatures. J. Phys.
Chem. 73, 34–39 (1969)
R.G. Larson, The Structure and Rheology of Complex Fluids (Oxford University Press, Inc,
New York, 1999)
D.R.E. Lide, CRC Handbook of Chemistry and Physics (CRC Press, Boca Raton, 2005)
V. Matoušek, A. Chryss, L. Pullum, Modelling the vertical concentration distributions of solids
suspended in a turbulent visco-plastic fluid. J. Hydrol. Hydromech 69, 255–262 (2021)
S.C. McCutcheon, J.L. Martin, J.R.J. Barnwell, Water quality, in Handbook of Hydrology, ed. by
D.R. Mainment, (McGraw-Hill, New York, 1993)
E.E. Michaelides, Particles, Bubbles and Drops. Their Motion, Heat and Mass Trransfer (World
Scientific Publishing, Singapore, 2006)
D.M. Murray, D.A. Holtum, Inter-conversion of Malvern and sieve size distributions. Miner. Eng.
9, 1263–1268 (1996)
I. Newton, Principia mathematica philosophiae naturalis. (1726)
J. Nikuradse, Strömungsgesetze in Rauhen Rohren (Laws of Turbulent Pipe Flow in Smooth Pipes).
VDI-Forschungsheft 361 (1933)
L. Prandtl, Neuere ergebnisse turbulenzforschung. VDI-Zeitschrift 77, 5 (1933)
L. Pullum, Similarity in suspension flows. (Keynote Address) Presented at the 20th Intl. Conf. on
the Hydraulic Transport of Solids in Pipes, Melbourne 2017 (2017)
D.V.S. Rao, Textbook of Mineral Processing (Scientific Publishers, Jodhpur, Rajasthan, 2017),
p. 2017
M. Rashidi, G. Hetsroni, S. Banerjee, Particles-turbulence interaction in a boundary layer.
Int. J. Multiphase Flow 16(6), 935–949 (1990)
J.F. Richardson, W.N. Zaki, Sedimentation and fluidisation: Part 1. Trans. Inst. Chem. Eng. 32,
S82–S100 (1954)
P.N. Rowe, A convenient empirical equation for estimation of the Richardson-Zaki exponent.
Chem. Eng. Sci. 42, 2795–2796 (1987)
N. Shillabeer, B. Hart, A.M. Riddle, The use of a mathematical model to compare particle size data
derived by dry-sieving and laser analysis. Estuar. Coast. Shelf Sci. 35, 105–111 (1992)
V.L. Streeter, E.B. Wylie, Fluid Mechanics (McGraw-Hill, New York, 1975)
P.K. Swamee, A.K. Jain, Explicit equations for pipe-flow problems. J. Hydr. Div. ASCE 102,
657–677 (1976)
A.M. Talmon, M. Huisman, Fall velocity of particles in shear flow of drilling fluids. Tunn. Undergr.
Space Technol. 20, 193–201 (2005)
L.W. Tilton, J.K. Taylor, Accurate representation of the refractivity and density of distilled water as
a function of temperature. J. Res. NBS 18, 205–214 (1937)
R. Turton, O. Leverspiel, A short note on the drag correlation for spheres. Powder Technol. 47,
83–86 (1986)
H. Wadell, Volume, shape, and roundness of quartz particles. J. Geol. 43, 250–280 (1935)
B.A. Wills, J. Finch, Wills Mineral Processing Technology (Butterworth-Heinemann, Oxford,
2015), p. 512
Wilson KC, Horsley RR (2004) Calculating fall velocities in non-Newtonian (and Newtonian)
fluids: a new view. Presented at the 16th Intl. Conf. on the Hydraulic Transport of Solids,
Cranfield, UK 2004
K.C. Wilson, R.R. Horsley, T. Kealy, J.A. Reizes, M. Horsley, Direct prediction of fall velocities in
non-Newtonian materials. Int. J. Miner. Process. 71, 17–30 (2003)
Chapter 3
Principles and Classification of Slurry Flow
3.1 Introduction
The old fable of the blind men and the elephant (Saxe 1872) was introduced in
Chap. 1. In the fable, each blind man had touched a different portion of the animal,
and they subsequently found themselves in greater disagreement than before. Slurry
flow, like an elephant, can be visualized in many ways, with results that sometimes
bring to mind the acrimonious deliberations of the blind men. To maintain a holistic
approach, it is necessary first to examine various aspects of slurry flow and then see
how they come together to form the “whole elephant.”
In general terms, slurries are mixtures of solid particles in a carrier fluid. The
carrier fluid may be water or some other Newtonian liquid, or it may be a
non-Newtonian liquid, or it may be a mixture of water and small particles that
behaves like a non-Newtonian fluid. This chapter provides a framework for classi-
fying the different regimes of slurry flow. Some of these regimes are common, while
others are unusual but serve as important exceptions. We will examine their physical
natures, the conditions under which they occur, and the key parameters affecting
their behaviors, many of which are developed in detail throughout the book.
This overview is important for several reasons. First, the variations are not
immediately obvious or easily correlated to our everyday experience; some dynam-
ics accord to common sense, but surprises will be encountered as well. Second, the
modes vary widely, both in their physical behavior and in the parameters governing
that behavior. Different modes often require different types of models, so that
success in modeling entails a proper identification of the mode at hand. Finally,
some modes may transition from one to the other as the slurry properties change, or
they may combine to form “hybrid” modes. An understanding of the basic governing
parameters is essential to recognize the more unusual combinations that may occur in
Before getting into the main text of this section, we first want to make a few
comments about how solids concentrations are expressed in various industries. As
you will discover, the amount of solids in a slurry generally has a major influence on
the slurry’s behavior, modifying the ways in which it flows, settles, and consolidates.
Industry is usually concerned with the mass of materials being transported or
produced; as a result, a concentration based on the weight of the solids and the
weight of the conveying material is often the preferred choice. As explained below,
however, a concentration based on the volume of the various phases provides a
clearer understanding of how the slurry will behave. For example, while two
different slurries made from different density materials with the same weight-
based concentrations are unlikely to behave in a similar manner, those with the
same volume-based concentrations generally will.
Concentrations based purely on weight or volume are not the only ones used.
Some industries state the ratio of the weight of solids to the weight of the fluid—a
simpler measure of concentration, but potentially misleading in application. Others
use the weight of solids in a given volume of fluid—a property that is no longer
dimensionless. Another common variation is the concentration of the fluid instead of
that of the solids. These pragmatic formulations may be useful on a day-to-day basis,
but they do little to further the understanding of the science.
For consistency and for the sake of effective modeling, the concentration of solids
throughout this book is expressed in volumetric terms, that is, as the ratio of the total
volume of solids to the total volume of the slurry mixture. Concentration is desig-
nated as Cv for the space-averaged volumetric concentration and as cv for the local
volumetric concentration at some point. Also, as we discuss in Chap. 6, if there is a
very high solids concentration, the simple volumetric value may no longer suffice to
describe the flow behavior; in these cases, an alternative, volume-based concentra-
tion, normalized with the maximum concentration obtainable for the solids in
question, may be used instead.
3.2 Properties of Slurry and Slurry Flow 59
When we turn from the flow of a simple liquid to that of a slurry (a two-phase
mixture of solid particles in a carrier fluid), we immediately need additional param-
eters to describe the properties of both the slurry and the slurry flow, as well as a
more precise system of nomenclature. For example, we must now distinguish
between several densities, including that of the fluid (ρf), that of the solid particles
(ρs), and that of the mixture (ρm). For many slurries, the carrier fluid is water, with a
density of approximately 1000 kg/m3 (our standard water; see Sect. 2.4). We
employ the symbol ρw for standard water (water at 5 °C) and ρf for the densities
of other fluids. The value of ρw forms the basis for expressing the relative density or
specific gravity of other solids or fluids. For example, there is a wide range of
applications where the solids being conveyed have a density of around
2650 kg/m3, which is a typical value for sand. In these cases, the relative density
of the solids is Ss = ρs /ρw = 2.65. Although the fluid is often standard-density water,
this is not always the case. For example, in marine dredging operations, the carrier
fluid is sea water, for which the relative density, denoted Sf, varies from place to
place, but has a typical value of about 1.03.
For a mixture of solids and fluid, the relative density or specific gravity of the
mixture is Sm = ρm/ρw, which is given by the general formula:
Sm = Sf þ ðSs- Sf Þ Cv ð3:1Þ
In this formula, Cv is the volumetric concentration (i.e., the fraction of the mixture
volume that is occupied by the solids), and the specific gravity of the carrying fluid
Sf = ρf /ρw. When the fluid is water of standard density (1000 kg/m3), Sf is unity and
the equation for Sm becomes Sm = 1 + (Ss - 1) Cv.
As noted previously, some industries also commonly use the weight concentra-
tion (Cw). Table 3.1 shows the relationships between the three common measures of
slurry concentration and density (Cv, Cw, and Sm). Figure 3.1 shows these same
relationships graphically for Sf = 1.
In specifying concentrations, be aware that delivered and in situ values represent
two different concepts of concentration. The in situ volumetric concentration (Cvi) is
the ratio of the volume of solids to the total volume of mixture within a pipe, while
60 3 Principles and Classification of Slurry Flow
1.0 6.0
0.9 Sm Cv S s 1 Cv 5.5
S Ss = 4.5
0.8 Cw Cv s 5.0
Sm 3.5
0.7 2.5 4.5
0.5 3.5
Sm
0.2 2.5 2.0
0.0 1.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6
Cv
Fig. 3.1 Relationships between volumetric concentration (Cv), weight concentration (Cw), and
specific gravity of slurry (Sm) for Sf = 1. Curves are for various values of specific gravity of particles
(Ss). Upper curves for Cw, lower curves for Sm
the delivered volume concentration (Cvd) is the ratio of the solids volumetric flow
rate to the total mixture flow rate moving through a pipe (or other flow conduits).
More specifically, the in situ concentration refers to the instantaneous value of
concentration at a specific location within the flow. It may be defined at a point, or as
an average value over a volume of slurry in a pipe. For example, if a length of the
pipe is isolated by suddenly closing high-speed valves at both ends (ignoring the
effects of water hammer), then the volumetric fraction of the solids in the isolated
pipe represents the in situ, or resident, value. More commonly, the in situ value is
defined as the average value over a two-dimensional section cutting across the
direction of flow.
In contrast, the delivered concentration refers to the average ratio of delivered
solids to the volume of delivered mixture. As a ratio of flow rates, this concentration
must be evaluated as the flux through a two-dimensional section cutting across the
slurry flow. It is often defined at the point where solids are delivered from (or fed to)
the conveying system. For example, if the slurry discharged from a system is
collected in a tank, then the volume fraction of solids in this tank equals the delivered
volumetric concentration. In a length of closed piping under steady-state conditions
where slurry is not being added or removed, the delivered concentration will remain
constant in order to satisfy conservation of mass; however, the in situ concentration
within the same pipeline may vary from section to section along the pipe.
Although it might initially appear that these two measures of concentration
should give the same value, this is only the case for truly homogeneous slurries in
which the solid and liquid components do not tend to segregate, slip, or otherwise
move relative to each other. The values of Cvd and Cvi differ when the average
3.2 Properties of Slurry and Slurry Flow 61
Vm LOW
Cvi > 0 Qs = 0
Cvd = 0
velocity of the solid particles (Vsp) is not the same as that of the fluid (Vf); and this is
typically the case in stratified flows (see Chap. 4). Although it is not necessary to
have stationary solids in the pipe for Cvd and Cvi to differ, an extreme example is a
pipe in which a deposit of stationary solids fills the lower half and water flows
through the upper half but moves only a few solid particles, which tend to roll along
the top of the bed. The in situ concentration Cvi is quite large, but as most of the
solids are not moving at all, the average velocity of the solids is small, and much less
than that of the water. As a result, the concentration of solids in the delivered mixture
(Cvd) would be much smaller than the Cvi in this case (see Fig. 3.2).
To find a formula that relates the two volume concentrations to each other, we
must express the flow rates of the two phases comprising the slurry. The volumetric
flow rate of liquid (Qf) is the product of Vf and the cross-sectional area occupied by
the fluid, i.e., Vf (1 - Cvi)πD2/4. Similarly, the volumetric flow rate of solids (Qs) is
VspCviπD2/4. The total flow rate of the mixture (Qm) is given by the sum of the fluid
and solids flow rates and is also equal to πD2/4 times the mean velocity of the
mixture (Vm). Thus:
4Qm
= V m = V f ð1- C vi Þ þ V sp Cvi ð3:2Þ
πD2
Qs V sp
C vd = = C ð3:3Þ
Qm V m vi
Equation (3.3) shows directly that the delivered concentration must be less than
the in situ value, provided Vsp is less than Vm. This condition is described as lag or
hold-up (or, less accurately, slip) of the solids. The lag or slip is the velocity
difference (Vm – Vsp), and the lag ratio (Λ) is obtained by dividing this quantity by
the mean velocity:
62 3 Principles and Classification of Slurry Flow
V m - V sp
Λ= ð3:4Þ
Vm
1
C vi = C ð3:5Þ
1 - Λ vd
Thus, when there is hold-up of the solids relative to the liquid, the in situ
concentration is greater than the delivered concentration, and the difference between
the two concentrations increases when the lag ratio increases.
This conclusion has a number of far-reaching implications. In particular, the
measurement of in situ solids concentration (by a radiation technique, e.g.; see
Chap. 12) does not directly represent the delivered concentration, which is often of
primary importance to the overall process being controlled or modeled. A further
corollary concerns the analysis of heterogeneous slurry transport, for which the hold-
up effect is significant. As will be seen, the in situ concentration is most important in
determining friction losses. However, process design methods are often concerned
with the delivered concentration, and the in situ concentration is either inferred or not
estimated explicitly. Furthermore, slurry testing often uses closed-loop systems, in
which the solids inventory and therefore the resident concentration is fixed; thus, the
delivered concentration will vary as Vm is changed. Hence, the simple approach,
suitable as it is for single-phase fluids, cannot be applied to slurries.
The pressure gradient for slurry flows, like that for single-fluid flows, can be
expressed in terms of the flowing medium (the hydraulic gradient j) or in terms of
the standard-density water (the gradient i). Thus, the pressure gradient (-dp/dL)m
associated with the flow of slurry in a pipe is expressed in the non-dimensional form
in one of two ways:
• As the hydraulic gradient based on the mean density of the slurry (ρm) (and the
associated relative density (Sm)):
1 dp 1 dp
jm = - = - ð3:6Þ
ρm g dL m Sm ρw g dL m
1 dp
im = - ð3:7Þ
ρw g dL m
The hydraulic gradient j is more useful for some purposes, such as matching
pipeline system and pump characteristics (see Chap. 10), since in many cases, the
head produced by a centrifugal pump is relatively independent of the mixture
density. Some slurries also exhibit hydraulic gradients that are only mildly affected
by mixture density when expressed in terms of j.
Note the similarity between Eq. (3.6) and the pressure term in the Bernoulli
Equation, i.e., ( p/ρg) from Eq. (2.16). Once jm is known for any slurry and pipe
diameter, the contribution of the hydraulic gradient to the total pipeline head is
simply equal to jmL, where L is the length of the pipeline section being considered.
The dimensionless pressure gradient i, on the other hand, is useful when an
engineer wishes to use a measure that is directly proportional to the pressure
gradient, regardless of mixture density. It is also useful in evaluating systems
where the mixture density varies from point to point. Much of the mathematical
modeling of slurry pressure gradients is done in terms of i. From Eqs. (3.6) and (3.7),
the two gradients are related by
im = Sm jm or jm = im =Sm ð3:8Þ
The density of slurry in Eqs. (3.6) and (3.8) can be based on either the in situ
concentration (Cvi) or the delivered concentration (Cvd). The latter is used more often
in the slurry flow models discussed in this book.
In the modeling of slurry flows, im is often expressed as the sum of the dimen-
sionless pressure gradient of the carrier fluid alone (if), plus an additional term
(or terms) for the effects of the solids, where if represents the gradient of the carrying
fluid expressed in terms of the height of water at standard density (ρw). Note that the
carrier fluid for a slurry model may be chosen to include some of the finer solids in its
composition and may differ from standard water in both density and viscosity.
The term iw is used for the pressure gradient of standard water. It may replace if in
a slurry model when the carrying fluid is closely approximated by water at standard
conditions (see Sect. 2.4).
Just as there are several elephants in a herd, many different slurries are conveyed in
the various industries. In the mining and minerals industry, slurries vary from low
concentration, finely divided solids in flotation tanks, cyclone separators, and pro-
cess feeds, to high concentration thickener underflows and thick viscous pastes used
for central discharge waste disposal. Particle sizes may be very carefully graded for
long-distance transport, very coarse for hydrotransport or hoisting operations, or
64 3 Principles and Classification of Slurry Flow
bi-modal for suspensions of fine tailings together with coarse rock in co-disposal
systems. The power industry uses low concentration suspensions when scavenging
electrostatic precipitators while pumping the waste ash out for disposal as a high
concentration, non-Newtonian viscous mixture. In chemical and process plants,
slurries of all forms abound, and in the food industry, slurries can vary from peas
in water to emulsified mixtures in mayonnaise and spreads. And then there are the
ubiquitous sludges processed at sewage treatment plants.
Clearly, slurries are not simple to classify, yet classification is essential if we are
to design effective slurry transport processes. So, accepting that some slurries will
have unique peculiarities that separate them from the rest, this section introduces
criteria to broadly classify slurries into various groups based on their multi-phase
fluid dynamics.
Most slurries can be broadly classified into one of two categories: settling or non-
settling. This classification is not sharp, and some slurries fall somewhere between
the two. Some may even transition from one to the other if their solids size or
concentration varies, or if the slurry undergoes a physical or chemical processing.
Within each category, a slurry can behave in a number of different modes depending
on the size, density, concentration, and other properties of the particles; the density,
viscosity, and other properties of the carrier fluid; and the velocity of the slurry
mixture within the pipe or channel.
A settling slurry is one in which most of the solids will settle to the bottom of a
container within an hour or so, if allowed to sit undisturbed. When the slurry is
flowing, the solids are supported by turbulence within the flow, or they find support
against the walls of the conduit that contain the flow, or there is some combination of
these. It usually is not possible to transport the solids in a laminar flow regime. The
transport medium is often a low viscosity Newtonian fluid, in many cases water. In
some ways, settling slurries are more straightforward than non-settling slurries, as
the parameters affecting their behavior are often less varied. On the other hand, their
behavior may vary dramatically with velocity and be subject to flow instabilities
within certain velocity ranges where the particles can easily segregate.
In a non-settling slurry, the particles do not readily settle except with the passage of
longer periods of time, in some cases, years. Settling may also occur with the
3.3 Classification of Slurry Mixtures and Flows 65
application of intense vibration or fluid shear, or with other treatments that suppress
particle-to-particle interactions or cause the smaller particles to agglomerate into
larger ones. Exceptions aside, in the actual practice of handling them, they behave as
if the particles do not settle.
Non-settling slurries are sometimes called non-Newtonian slurries, though both
settling and non-settling slurries can behave in ways that are not Newtonian. Still,
non-settling slurries do often have carrier liquids that are highly viscous and exhibit
non-Newtonian characteristics. In many cases, the finer particles within a
non-settling slurry will interact strongly with each other to create what is essentially
a pseudo-fluid with non-Newtonian properties. Such interactions may occur due to
chemical or electrostatic effects (such as with clays), or from physical interactions
due to a high concentration of solids. A settling slurry, especially one with finer
particles, may transition to a non-settling slurry and behave like a non-Newtonian
pseudo-fluid if the concentration becomes high enough for the particle-to-particle
interactions to play a dominant role in the slurry flow. Because the viscous effects of
the fluid (or pseudo-fluid) are often strong, non-settling slurries may often transport
solids in both the laminar and turbulent flow regimes. In some ways, non-settling
slurries are more complex and varied than settling slurries; however, they may be
less subject to instabilities due to their reduced tendency to experience particle
segregation.
A slurry of very fine particles (say less than 40 μm, a conventional division between
silt and sand sizes) tends to behave in a homogeneous fashion, where the particle
distribution is virtually uniform throughout the flow. In these cases, there is almost
no change of solids concentration (cv), with height within the pipe or flow conduit
(see cv profile in Fig. 3.3a). The distribution of particle velocities (us) is axisymmet-
ric across the pipe cross section (see us profile in Fig. 3.3a), and there is no local lag
between the velocity of the particles and that of the carrying fluid. Therefore,
Cvd = Cvi. Fine-particle settling slurries with Newtonian carrier liquids must usually
operate in turbulent flow to behave as a homogeneous slurry, but the velocity at
which the particles begin to settle will be very low. Non-settling slurries (particularly
if the concentration of fine particles is high, or if the particles are very rheologically
active) may support their solids without turbulence and operate in either a laminar or
turbulent flow regime. In both settling and non-settling slurries, the mixture of the
liquid with fine particles may often be treated as a pseudo-fluid or carrier fluid with
properties determined by the rheological interaction between the particles and the
fluid.
A settling slurry of somewhat larger particles (roughly between 40 and 200 μm,
which is the fine sand range) exhibits slightly different behavior. The particles are
rheologically inactive, and the flow will approximate Newtonian behavior if the
carrier liquid is itself Newtonian. The particles may separate from the carrying liquid
66 3 Principles and Classification of Slurry Flow
at very low velocities, but at most velocities typical of industrial application, all
particles will be supported by turbulent eddies. There will be a weak concentration
gradient and a measurable decrease of concentration with increasing height in the
pipe cross section (Fig. 3.3b). The corresponding velocity distribution will be
slightly asymmetrical (Fig. 3.3b), and the solids slip (lag) will be negligible. This
type of flow regime is known as pseudo-homogenous flow.
If the size of transported particles is in the medium sand range (typically between
200 and about 800 μm), the settling slurry exhibits heterogeneous behavior, with
smaller values of Vm producing greater non-uniformity of distributions of both solids
concentration and solids velocity (Fig. 3.3c). For grains of sizes between about
800 μm and 1.5% of the pipe diameter (typical coarse sand and fine gravel), the
heterogeneous flow becomes more distinctly stratified, with a detectable sliding bed
(particularly in highly concentrated slurries). The transported grains can be
supported above the sliding bed by turbulent eddies, mutual collisions, or both.
This partially stratified flow has profiles of both local concentration and velocity
composed of two distinct parts: an area of approximately uniform distribution across
the sliding bed, and an area of steep gradient in the profiles just above the bed
(Fig. 3.3d). The pipe cross-section average slip is significant; thus, Cvd < Cvi.
Fully stratified flow occurs with settling slurries if virtually all of the transported
particles occupy the bed, and turbulent suspension of particles is ineffective. The
ratio of particle size (d) to the pipe diameter (D) is of major importance in deter-
mining the presence of this flow type, which does not normally occur for d/D ratios
less than 0.015. The cv profile (Fig. 3.3e) exhibits a sharp gradient at the interface
between the bed and the carrier flow above the bed, and this gradient also affects the
local velocity distribution. The area of the sharp gradient is associated with the part
3.4 Physical Mechanisms of Particle Support and Friction 67
of the bed sheared by the flow and it is called a shear layer. The average slip across
the pipe cross section tends to be significant and typically Cvd < < Cvi. Fully
stratified flow is less likely to occur if the particles are broadly graded (i.e., spanning
a wide range of particle sizes).
In a settling slurry with a viscous liquid (i.e., more than 20 times that of water),
the above dynamics may change, as viscosity plays a larger role in both the support
of the particles and the onset of turbulence. However, such slurries are rare, perhaps
excepting certain applications in the process and chemical industries, and they are
not treated with any detail in this text. More common are non-settling slurries with
non-Newtonian carrier fluids, often consisting of water mixed with very fine parti-
cles. When these slurries also contain coarse particles that can segregate within the
flow, they are referred to as complex slurries. In laminar pipe flow, the coarse solids
within these slurries are often transported as a fully stratified bed (Chap. 6), but
the flow dynamics of such slurries differ from their settling slurry counterparts
mentioned above, and they appear to superficially behave as homogeneous
non-Newtonian fluids when we examine their pressure gradient versus flow curves.
The uncertainty in determining the modes of particle support in these various
instances of slurry flow can cause significant difficulty in interpreting flow behavior
in a pump-pipeline system. Note also that while a particle size may fall into one of
the above ranges/flow regimes, it may exhibit another flow regime at extreme ends of
the velocity spectrum. For example, a slurry with a heterogeneous particle size
distribution can become fully stratified at low velocities, or pseudo-homogenous at
very high velocities (see Sect. 3.6). However, the flow regimes and associated
particle sizes discussed above tend to hold true within the industrially relevant
range of flow velocities for a given pipe size and are thus useful, in this respect,
from a modeling perspective.
When flow is fully stratified, the particles are concentrated in the lower portion of the
pipe and will come into contact with each other and with the pipe wall. This contact
can be continuous, as occurs with a stationary or sliding bed of solids, or it can be
sporadic, as occurs when the travelling particles are partially supported by turbulent
eddies, but also by intermittent collisions with each other and the pipe wall.
The submerged weight of particles carried by granular contact produces normal
granular stresses against the bottom of the pipe. Shear stresses proportional to the
normal stresses are necessary to set the particles in motion. In the case of continuous
contact, the stresses are covered by the friction concepts of Coulomb (1773) and
hence are denoted as Coulombic solids stresses. The solids stresses in a sheared
granular body of sporadically colliding particles are evaluated by Bagnold’s concept
(Bagnold 1954; Bagnold 1956) and are thus called Bagnold solids stresses.
Bagnold (1956) also considered the total solids concentration in any unit volume
to consist of elements of both bed load (also called contact load) and suspended load,
with the submerged weight of the bed load transmitted downward through the
intergranular normal stress, and that of the suspended load “transmitted not to the
bed grains but between them as an excess static fluid pressure.” Hence, the sub-
merged weight of the suspended particles was transmitted to the carrying liquid,
effectively increasing its density.
This concept of the two mechanisms of particle support (fluid action and
intergranular contact) provides the basis for understanding the behavior of slurry
flows. A further elaboration of this concept for the pipe flow of stratified settling
slurry, published by Wilson and his co-workers (e.g., Wilson 1970; Wilson 1987;
Wilson and Pugh 1988; Pugh and Wilson 1999) provided a detailed description of
solids stresses across flow layers (bed layer and shear layer) at their interfaces and at
the flow boundary (usually pipe wall).
A detailed analysis of the solids stresses is presented in Chap. 4, where they are
used in a predictive two-layer model for stratified flows.
If the particles are smaller than the contact-load particles, they will be suspended by
fluid turbulence and are thus known as suspended load. All transported particles
contribute to the suspended load in pseudo-homogeneous flow, while in heteroge-
neous flow some portion of the particles contribute to the suspended load and the rest
of them contribute to the contact load.
Turbulent diffusion produces characteristic shapes of concentration profiles in
slurry pipes; those can be found in publications presenting measurement results from
several slurry test laboratories (e.g., CSIRO, Czech Academy of Sciences, Delft
University of Technology, IITD, Saskatchewan Research Council). The concept of
3.4 Physical Mechanisms of Particle Support and Friction 69
Some slurry-flow experiments have revealed effects that can be attributed to the
presence of near-wall lift that repels particles from the lowermost portion of the
boundary layer and prevents their contact with a pipe wall, hence diminishing the
flow friction. The presence of a significant off-wall lift force was indicated as most
important for particles in a size range substantially larger than the thickness of the
viscous sublayer. It occurs in both coarse pseudo-homogeneous flows (particles
larger than around 90 μm) and heterogeneous flows. Analyses and their experimental
support are available in various published studies, e.g., Wilson et al. (2000), Wilson
and Sellgren (2003), Whitlock et al. (2004), and Wilson et al. (2010).
70 3 Principles and Classification of Slurry Flow
In a horizontal straight pipe of constant diameter, the drop of static pressure in the
slurry flow along the length of the pipe is due exclusively to friction. The pressure
drop per unit length of pipe, Δp/L (i.e., the pressure gradient), is a measure of the
friction loss, and it is often expressed in units of the head as the hydraulic gradient
( jm) or the dimensionless pressure gradient (im) (see Eqs. 3.6 and 3.7). This gradient
varies with the flow velocity (Vm), and a curve showing the relation between im and
Vm (or the relation between jm and Vm) is called the pipe characteristic curve.
Different types of slurries exhibit different shapes of pipe characteristic curves
within the range of flow velocities of practical interest. Several characteristic thresh-
old values of Vm are of interest, some of which are associated with the characteristic
shape of the pipe curve.
Figure 3.4 displays schematic shapes of pipe characteristic curves for different
types of slurries at constant Cvd, as compared with the curve for flow of water
obtained by the Darcy-Weisbach formula (see Eq. 2.46).
a) b)
Pressure gradient
Pressure gradient
Cvd or rm
Ss or rs
Pressure gradient
c) d)
Cvd or rm
typical
centrifugal
pump pseudo-
characteristic sliding bed
homogeneous
heterogeneous
sliding bed
sliding bed
stationary bed stationary bed
Velocity, flow rate Velocity, flow rate
Fig. 3.4 Schematic pipe characteristic curves at constant Cvd for different types of slurry flow and
internal structure of different types of slurry flow: (a) non-Newtonian homogeneous flow,
(b) pseudo-homogeneous flow, (c) heterogeneous flow, (d) fully stratified flow
3.5 Friction Loss and Pipe Characteristic Curve 71
The curves in Fig. 3.4a have shapes typical for the non-Newtonian homogeneous
slurries that are common in the mining and mineral industry. In these slurries, the
fine solids are rheologically active and significantly alter the rheological parameters
of the carrying liquid. The shapes of the curves depend upon the slurry’s rheology, a
complex function of particle size (d ), shape, surface chemical properties, fluids
chemical properties, temperature (T ), pH, and, in particular, solids concentration
(Cv), as shown.
Pseudo-homogeneous settling slurry flows have characteristic curves that are
similar in form to those of their Newtonian carrier fluids (often water), exhibiting
a pressure gradient that increases geometrically with throughput velocity in the range
of flow velocities where all particles are suspended (Fig. 3.4b). The variation in
frictional head loss of the pseudo-homogeneous flow at velocities of full suspension,
therefore, follows a fluid-like form, and particles contribute to flow friction through
their interaction with the carrying liquid rather than through contacts (collisions)
with each other or with a flow boundary (pipe wall). The curves become steeper with
increasing delivered concentration of solids (and thus mixture density, or ρm) or
higher density of solids (ρs) (and, thus, relative density of solids, or Ss), indicating
that the increase in these parameters increases the friction losses in a pipe.
In heterogeneous flow, pipe curves for slurry flows of a constant, delivered
concentration exhibit minimum points, as illustrated by Fig. 3.4c. The upward
trend of im (or Δp/L), with decreasing mixture velocity Vm below the velocity at
which im is minimal, is associated with increasing stratification and the establishment
of a granular bed at the bottom of the pipe. This development, along with related
threshold velocities, is discussed in Sect. 3.5. The minimum in the pipe curve has
consequences for its interaction with a typical pump characteristic curve that will be
analyzed in Chap. 10. Note that if the flow velocity increases, then the pipe curve
tends towards the curve of the carrying liquid, indicating that the contribution of
solids to the total pressure gradient decreases with the increasing Vm. This change
is due to the decreasing proportion of contact-load particles (and the increasing
proportion of suspended-load particles) that accompany the increase of Vm in a
heterogeneous flow.
As discussed before, contact-load particles cause significantly more resistance
than suspended particles. At very high velocities, the particles in a heterogeneous
flow may become fully suspended, and the flow transform to pseudo-homogeneous
flow. However, the required velocity may be well above that encountered in practical
operations. Conversely, at very low velocities, the flow may reach the fully stratified
condition where no particles are in suspension, but again, this may happen at a
velocity that is outside the practical range of normal operations.
When very coarse particles with density greater than the carrying fluid are
transported, fully stratified flow is the only plausible flow mode. Such particles are
too large and heavy to be suspended at any realistic flow velocity in a pipe. Hence,
contact friction is the exclusive mechanism through which the transported solids
contribute to the pressure gradient in the entire range of potential operating veloc-
ities. The contact distribution is weakly sensitive to the flow velocity; therefore, the
pipe characteristic curve of the fully stratified flow does not tend to the carrying
72 3 Principles and Classification of Slurry Flow
liquid curve (Fig. 3.4d). Although fully stratified flow is less energy efficient than
heterogeneous flow, it can be economically attractive for pipelines of moderate
length, as it can remove the need to reduce the particle size before pumping, an
operation that itself requires considerable energy. Avoiding the presence of finer
solids may also reduce the cost and environmental impact of material separation after
pumping.
In evaluating the extra friction loss caused by the conveyed solids (i.e., the solids
contribution to the total pressure gradient), we refer to the solids effect (im - if),
where if is the friction gradient for the carrier liquid alone at any given mixture
velocity (Vm). The evaluation of this quantity is discussed extensively in Chaps. 4, 5,
and 6.
The characteristic velocities of slurry flow are specific reference and/or threshold
values of the mixture velocity Vm. They are employed to delimit ranges of flow for
different forms of particle support, different flow patterns, and different shapes of
pipe characteristic curves (Fig. 3.5).
The characteristic velocity (Vs) is the deposition limit velocity. In other words, it is
the flow velocity at which a stationary deposit starts to be formed in slurry flow, as
the first particles stop moving and begin to accumulate at the bottom of the pipe. This
velocity is an important parameter in an evaluation of flow economy, safety, and
stability. It occurs in all types of settling slurry flows, although its value differs
significantly with particle size, particle density, and pipe size. The deposition limit
velocity is discussed in detail in Chap. 4. Furthermore, in Chap. 5, we discuss its use
in a model to predict friction loss in fully stratified flow.
The minimum velocity (Vmin) is the flow velocity (Vm) at which a pipe character-
istic curve exhibits a minimum in friction loss. This velocity is important in
heterogeneous and stratified flows for reasons associated with flow stability, as
discussed in Chap. 10. Its value is often close to that of the deposition limit velocity
(Vs), although the two cannot be considered the same. For instance, Vmin often
increases with increasing Cvd in a pipe, while the deposition limit velocity tends to
decrease with increasing Cvd. The employment of Vmin to the description of stratified
flows is further discussed in Chap. 5.
The velocity at incipient turbulent suspension (V0) is the threshold Vm at which
the first particles start to be picked up and suspended by turbulent eddies of carrying
liquid in slurry flow. Below this velocity, all particles are transported as contact load,
and the flow is fully stratified.
The threshold velocity V100 is the flow velocity at full (100%) suspension. Above
this velocity, all particles are transported as suspended load. Hence, it is the upper-
limit velocity for the heterogeneous flow and the transitional velocity to the pseudo-
homogeneous flow regime. One way to determine this threshold velocity is to
consider it as equal to Vm at the intersection of the im-Vm pipe characteristic curves
for the pseudo-homogeneous flow and the heterogeneous flow (Newitt et al. 1955;
Matoušek et al. 2018).
The characteristic velocity V50 represents the flow velocity at which one-half
(50%) of grains are suspended by the carrier turbulence, and the other half is
transported as contact load. In Chap. 5, we discuss this velocity’s use in a model
to predict friction loss in heterogeneous flow.
Be aware that most pipe system curves for settling slurry flows do not exhibit the
entire sequence of characteristic velocities as schematically shown in Fig. 3.5. For
example, the deposition limit velocity may be higher than the incipient suspension
velocity (Vs > V0), which is typical for heterogeneous flows of weak stratification;
hence, V0 disappears from the velocity range of practical interest covered by the
pipeline system curve. On the other hand, pseudo-homogeneous settling slurries
exhibit deposit formation at very low Vm, and their V100 tends to be only slightly
larger than Vs.
The practical usefulness of the threshold velocities Vs, Vmin, V0, and V100 becomes
obvious in Chap. 5, where modeling of different settling slurry flows is discussed.
Figure 3.6 also demonstrates the usefulness of the characteristic velocities Vs and
Vmin. In flows exhibiting minimum points on pipe curves as illustrated in Figs. 3.4
and 3.5, it is of interest to consider the course of the curve at flow velocities
Vm < Vmin. In this velocity range, the difference between Cvd and Cvi increases
considerably with decreasing Vm as the depth of the bed of deposited solids
increases. This increase has a profound effect on the shape of the curve. Usually,
Vmin is close to the deposition limit velocity Vs, and there are even slurry flows where
Vmin < Vs, both in the field and laboratory.
For many industrial pipelines, the constant delivery condition is typical (e.g., Cvd
is constant at the system inlet during an operation). The in situ concentration (Cvi)
then develops based on the flow conditions. As Vm decreases and the difference
between the solids and fluid velocities increases, the associated Cvi also increases.
74 3 Principles and Classification of Slurry Flow
i i
(H) (H)
ipg
Vs V (Q) Vs V (Q)
Fig. 3.6 Comparison of schematic pipe characteristic curves of heterogeneous flow at constant Cvd
and at constant Cvi
This leads to increasing friction loss (im) as the volume of the sliding bed of
transporting particles grows, resulting in a pipe characteristic curve as shown in
the left-side plot of Fig. 3.6. As friction losses increase, the velocity will further
decrease unless immediate action is taken to increase the head driving the system
(such as by increasing pump speed). However, assuming that pump speed remains
constant or is not adjusted quickly enough, the trend of decreasing velocity and
increasing Cvi will continue until the entire cross section of the pipe is filled with
solids (i.e., plug flow) and the velocity approaches zero. It is theoretically possible
for flow to continue at this point, provided there is enough driving force in terms of
pumping power to overcome the plug-flow friction loss (iplug). In actual practice,
however, this power is usually not available, and the flow stops with a plugged
pipeline. In Chap. 5, we discuss the determination of iplug to determine the incipient
motion of a plug in a pipe.
For laboratory closed loops, the constant inventory solids condition is typical (Cvi
is approximately constant at various velocities in the loop) and Cvd decreases with
the decreasing Vm as grains settle out to form a deposit at Vm < Vs. The result is that
im tends toward zero at negligibly low Vm, as shown in the right-side panel of
Fig. 3.6. At low velocities below Vs and below V0, the curve represents the friction
of the carrying liquid only, as it moves through the cross-sectional area above the
deposit.
H
SEC = ð3:9Þ
Ss Cvd L
im
SEC = ð3:10Þ
Ss C vd
2.0
1.8 d50 103 Pm, DD==158
d50 =0.103mm, 158mm,
mm,V V
= m2 =m/s
2 m/s
SEC (kWh/tonne-km)
1.6 d
d50 = 103 Pm, D =158 mm, V
50 0.103mm, D = 158mm, V = m 4 =m/s
4 m/s
1.4 d = 370 Pm, D =150 mm, V
50 = 370mm, D = 150mm, V = 3
d50 m m/s m/s
= 3
1.2
1.0
0.8
0.6
0.4
0.2
0.0
0 0.1 0.2 0.3 0.4 0.5 0.6
Cvd (-)
Fig. 3.7 SEC versus Cv plots. Fine sand slurry
3.8 Case Studies 77
2.0
1.8 D = 203 mm, Vm = 3.3 m/s
SEC (kWh/tonne-km) 1.6
1.4
1.2
1.0
0.8
0.6
0.4 880Pm
d50 ==880
sand, d50 mm
0.2 granite rock
Granite d50 ==6100
rock,, d50 Pm
880mm
0.0
0 0.1 0.2 0.3 0.4
Cvd (-)
Fig. 3.8 SEC plots for two different aqueous slurries at Vm = 3.3 m/s in 203 mm pipe. Experi-
mental data from GIW Hydraulic Lab
larger pressure gradient and excessive values of SEC. The data for heterogeneous
flow of medium sand slurry (d50 = 370 μm) at a velocity not far above the deposition
limit velocity, collected in a lab circuit of the Delft University of Technology, show a
similar effect of the concentration on SEC (Matoušek 2002).
An example of experiment-based information on SEC for coarser slurries is given
in Fig. 3.8. This information is based on the results of GIW Hydraulic Lab tests
carried out for various aqueous settling slurries in a 203-mm pipe loop. Although the
test results do not include data for very high concentrations, they do confirm the
trend observed in the finer slurries at Cvd < 0.3. They also demonstrate that
obviously, coarser slurries consume more energy than finer slurries due to a higher
proportion of contact load contributing to flow friction.
Naturally, an accurate prediction of SEC and its sensitivity to the flow velocity
(Vm) and solids concentration (Cvd) depends on an accurate predictive model for the
frictional head loss in slurry flow (Hashemi et al. 2014). Predictive models for
frictional loss are introduced in Chap. 5 (for settling slurry flows) and Chap. 6 (for
non-settling slurry flows).
The piping system (Fig. 2.18) from Example 2.1 is found in a mineral processing
plant, transferring tailings to a tailings storage facility. Solid particles of the density
2650 kg/m3, with an approximate size between 70 and 160 μm, are mixed with water
(T = 20 °C) to form a Newtonian pseudo-homogeneous slurry in which the
78 3 Principles and Classification of Slurry Flow
volumetric concentration of solids is 0.25. The slurry flows at 0.12 m3/s (the same
total discharge as in Example 2.1) through the system’s 8-inch steel pipe of internal
diameter D = 0.2027 m and wall roughness ε = 45 μm.
Calculate the:
(a) Weight concentration of solids in slurry
(b) Density of slurry
(c) Discharge of solids
(d) Hydraulic gradient and pressure gradient
(e) Head required from the pump
(f) Specific energy consumption in the slurry pipeline
Ss C v 2:65 0:25
Cw = = = 0:469
Sf þ ðSs - Sf ÞC v 0:998 þ ð2:65 - 0:998Þ0:25
V 2m 0:0153ð3:72Þ2
jm = f =
2gD 2ð9:81Þð0:2027Þ
= 0:053 ðm head of slurry per m pipeÞ and from Eq: (3.8),
the dimensionless pressure gradient is
References 79
dp
- = im ρw g = 0:075ð1000Þð9:81Þ = 736 ðPa per m pipeÞ:
dL
X V2
L
H PUMP = Δz1 - 2 þ 1 þ f 1 - 2 þ K m
in m head of slurry: For
D 2g
V m = 3:72 m=s,
495 3:722
H PUMP = 35 þ 1 þ 0:0153 þ 0:9
0:2027 19:62
= 62:7 m of the slurry column:
Hw 88:5
SEC = = = 0:270 ð - Þ
Ss C vd L 2:65 0:25 495
References
R.A. Bagnold, Experiments on a gravity-free dispersion of large solid spheres in a Newtonian fluid
under shear. Proc Roy Soc A 225, 49–63 (1954)
R.A. Bagnold, The flow of cohesionless grains in fluids. Phil Trans Roy Soc A 249, 235–297 (1956)
C.A. Coulomb, Essai sur une application des règles de maximis et minimis à quelques problèmes de
dtatique, relatifs à l’architecture. Memoires de mathématique et de physique, l'Académie Royale
des Sciences 7, 343–382 (1773)
S.A. Hashemi, K.C. Wilson, R.S. Sanders, Specific energy consumption and optimum operating
condition for coarse-particle slurries. Powder Tech 262, 183–187 (2014)
80 3 Principles and Classification of Slurry Flow
V. Matoušek, Distribution and friction of particles in pipeline flow of sand water mixtures, in
Handbook of Conveying and Handling of Particulate Solids, ed. by H. Kalman, (Elsevier
Science, Amsterdam, 2001), pp. 465–471
V. Matoušek, The pipeline transport of different sand fractions in dense slurries. Paper presented at
the 3rd specialty conference on dredging and dredged material disposal, Orlando, FL, USA
2002 (2002)
V. Matoušek, R. Visintainer, J. Furlan, A. Sellgren, Threshold criteria for components of predictive
model for pipe flow of broadly-graded slurry. Presented at the ASME 5th Joint US-European
Fluids Eng. Summer Conf. FEDSM, Montreal, Canada 2018 (2018)
D.M. Newitt, J.F. Richardson, M. Abbot, R.B. Turtle, Hydraulic conveying of solids in horizontal
pipes. Trans. Inst. Chem. Eng. 33, 93–113 (1955)
F.J. Pugh, K.C. Wilson, Role of the interface in stratified slurry flow. Powder Tech 104, 221–226
(1999)
H. Rouse, Modern conceptions of the mechanics of fluid turbulence. Trans Amer Soc Civ Eng 102,
436–505 (1937)
J.G. Saxe, The blind men and the elephant. Fables and legends of many countries rendered in
rhyme, in The Home Book of Verse, ed. by Stevenson, (Holt, Reinhart and Winston, New York,
1872), pp. 1877–1879
C.A. Shook, M.C. Roco, Slurry flow: Principles and practice (Butterworth-Heinemann, Oxford,
1991)
W. Schmidt, Der massenaustausch in freier luft und werwandte erscheinungen. Die
Wasserwirtschaft No. 5–6 (1932)
R. Visintainer, J. Furlan, G. McCall, A. Sellgren, V. Matoušek, Comprehensive loop testing of a
broadly graded (4-component) slurry. Presented at 20th Intl. Conf. on Hydrotransport, Mel-
bourne 2017 (2017)
L. Whitlock, K.C. Wilson, A. Sellgren, Presented at 16th Intl. Conf. on the Hydraulic Transport of
Solids, Cranfield, UK 2004 (2004)
K.C. Wilson, Slip point of beds in solid-liquid pipeline flow. J Hydr Div ASCE 96, 1–12 (1970)
K.C. Wilson, Analysis of bed-load motion at high shear rates. J Hydr Eng ASCE 113, 97–103
(1987)
K.C. Wilson, Energy consumption for highly-concentrated particulate slurries. Presented at the 12th
Intl. Conf. on Transport and Sedimentation of Solid Particles, Prague, 2004 (2004)
K.C. Wilson, F.J. Pugh, Dispersive-force modelling of turbulent suspension in heterogeneous slurry
Flow. Can. J. Chem. Eng. 66, 721–727 (1988)
K.C. Wilson, A. Sellgren, Interaction of particles and near-wall lift in slurry pipelines. J Hydr Eng
ASCE 129, 73–76 (2003)
K.C. Wilson, A. Sellgren, G.R. Addie, Presented at the 10th Intl. Conf. on Transport and Sedimen-
tation of Solid Particles, Wrocław, Poland, 2000 (2000)
K.C. Wilson, R.S. Sanders, R.G. Gillies, C.A. Shook, Verification of the near-wall model for slurry
flow. Powder Tech 197, 247–253 (2010)
J. Wu, L. Graham, S. Wang, R. Parthasarathy, Energy efficient slurry holding and transport.
Minerals Eng 23, 705–712 (2010)
Chapter 4
Stratification of Slurry Flow and Deposition
of Solids in Pipes
4.1 Introduction
The previous chapter presented a survey of the basic mechanisms governing the
support and friction of solid particles in liquid flows. In the present chapter, these
mechanisms are further analyzed and employed in the modeling of stratified slurry
flow based on a layered structure of the flow. The main aim of this modeling is to
predict the frictional pressure gradient and the deposition limit velocity in a hori-
zontal pipe, although the analysis can be readily extended to inclined pipes discussed
in Chap. 7.
The analysis of stratified slurry flows deals primarily with the contact load and its
role in solids friction, as well as the processes of stratification and deposition in
settling slurry flows. These include fully stratified flows, in which coarse particles
travel entirely as contact load, and heterogeneous (partially stratified) flows, where
the contact load provides only partial support of the transported solids. The princi-
ples for a layered model of stratified flow are developed, and a survey of some major
predictive layered models is given.
The contact load comprises stationary beds, sliding beds, and colliding particles,
all of which tend to have a significant adverse effect on the pressure gradient. For the
stationary-bed case, problems of flow stability can arise; therefore, stationary
deposits are generally avoided in practice—although, as will be shown later, they
have an important role to play in modern tailings disposal techniques. In any case,
designers and operators must have a method for predicting stationary deposit
conditions, and this is one of the major benefits of a contact-load analysis facilitated
by a layered model. For finer settling slurry flows where the contact load is absent
(e.g., pseudo-homogeneous flows), alternative predictive models are suggested.
If a slurry contains particles large and massive enough to settle during transport
within a flowing liquid, the slurry will tend to stratify due to interactions between the
particles, the carrying liquid, and the flow boundaries. In the case of pipe flow,
various patterns develop, and as described in Chap. 3, the flow velocity is an
important parameter. If the flow stratifies so that at least a portion of the solids are
transported as contact load, then distinct layers can be established with different
types of granular contact. It was seen that intergranular contact can be continuous
(where the particles involved form stationary or moving beds) or sporadic (as when
particles collide with each other in the shear layer just above the upper surface of a
bed). If particles are fine and interact with the turbulence of the carrying liquid rather
than with each other, then a layer of turbulent suspension can develop above a
sliding bed of the same particles. If the flow velocity is not high enough to move all
of the particles, then a stationary deposit develops at the bottom of the pipe.
Figure 4.1 shows a dismantled industrial slurry pipe with more than half of its cross
section filled by a stationary deposit. Before dismantling, the pipe was operated as a
part of a tailings pipeline transport system. The stationary deposit built up over
several weeks and was caused by the introduction of high-density ore into the
tailings. The deposit buildup resulted in an effectively smaller pipe diameter and
correspondingly higher friction loss. This reduced the flow rate to approximately
one-half of the designed flow rate. It was the reduction in the flow rate that alerted the
Fig. 4.2 Stationary deposit in a laboratory pipe: (a) glass ballotini in water through a transparent
pipe; (b) sand distribution measured in the cross section of 150 mm pipe. (Matoušek 2007)
operators to the presence of the stationary bed in the pipe, which was otherwise not
visible at the pipeline outlet. Even with the stationary deposit, the pipe still
discharged some solids (both large and small) at its outlet. An explanation for this
observation will be given in Sect. 4.4.
In laboratory loops, for which the constant inventory of solids condition is typical
(see Sect. 3.6), it is relatively easy and safe to observe the development of a
stationary deposit and to measure conditions in the pipe flow above the deposit. In
fact, such observation is a prerequisite of many test programs.
Figure 4.2 shows the results of typical observations. In (a), a stationary deposit of
ballotini (fine glass beads) with the top of the bed sheared by a flow of water is
photographed in a transparent pipe. In (b), a measured solids distribution in the pipe
flow of a sand-water slurry identifies the presence of a deposit of up to 25% of the
flow depth and a shear layer extending through most of the remaining flow above the
deposit. The measurement confirms that the distribution of solids concentration is
virtually uniform in the deposit and is approximately linear with depth across the
shear layer.
As the name implies, a two-layer model considers stratified flow as flow composed
of two layers (lower and upper) that interact with each other and with a flow
boundary. In the case of fully stratified flow, solids are transported only in the
lower layer, and the flow through the upper layer is free of particles (Fig. 4.3).
84 4 Stratification of Slurry Flow and Deposition of Solids in Pipes
upper layer
(V1)
interface
a
D (W12 , f12)
b
lower layer
actual concentraon two-layer (V2 , Cvb)
distribuon model
Fig. 4.3 Schematic pattern of fully stratified flow for two-layer model
In partially stratified flow, the particles are also transported through the upper
layer. Figure 4.4 demonstrates the correspondence between a typical partially strat-
ified flow and its idealized characterization as a two-layer model. The virtual
interface between the two layers is defined as that depth in the flow where the
local gradient of solids concentration and velocity become steep, thus separating the
zone of high local concentration with a small local gradient at the bottom of the pipe
from the zone of gradually decreasing local concentration and larger gradient above
the interface.
4.3 Modeling of Stratified Flows 85
Key parameters of the two-layer model are considered separately for each layer,
identified by the subscripts “1” for the upper layer and “2” for the lower. These
include cross-sectional averaged values of the velocity (V1 and V2) and concentration
(C1 and C2), each defined over the corresponding cross-sectional areas (A1 and A2).
The local slip (hold-up) between phases is neglected within either of the layers.
Therefore, V1 is the average velocity for both liquid and solids in the invert above the
bed, and V2 is the velocity of both phases within the bed. In the fully stratified case
depicted in Fig. 4.3, C1 = 0. The perimeters are defined as O1 along the upper layer
boundary formed by the pipe wall, O2 along the lower layer at the pipe wall, and O12
along the width of the surface between layers. All these quantities are functions of
the position of the interface (β) and of the pipe diameter.
The model applies equations that express the conservation of mass for the flow in
the entire cross section per unit length of pipe and the conservation of momentum for
each of the two layers. The momentum equations for the layers are formulated as
force-balance equations. In these equations, some of the forces represent shear
stresses applied to the various boundaries of each layer, and others represent body
forces, such as the pressure gradient, acting within the flow itself. In developing this
force balance, it is appropriate to begin with the component that usually dominates
the resisting forces in the force balance, namely the solids force (and corresponding
stress) that the contact-load particles in the lower layer exert against the pipe wall
boundary.
Consider a pipe of diameter D with a bed of solids occupying the lower portion as
in Fig. 4.3. In the figure, α represents an angular coordinate defining the position of
any point on the pipe wall, and β gives the position of the upper surface of the bed.
The normal stress between the solid grains and the pipe wall is denoted σ s. The total
normal force per unit length, FN, is obtained by integrating the normal stress over the
boundary between the bed and the wall, giving:
Z β
FN = D σ s dα ð4:1Þ
o
The total frictional force between the sliding solid grains and the pipe boundary
equals μsFN, where μs is the coefficient of mechanical friction between the particles
and the wall material.
Note that the total normal force (FN) exerted by the bed solids is not equivalent to
the submerged weight of the solids. A summation of forces in the vertical direction
shows that only the vertical component of σ s can act to support the bed weight; thus,
the submerged bed weight (FW) will be inherently less than the expression for FN
found above, which includes both vertical and horizontal components (Wilson 1970,
Wilson et al. 1972).
A relationship for the variation of σ s with depth is based on the simple assumption
(Wilson 1970) that the rate of increase of granular pressure with depth is a constant
and equals the submerged unit weight of solids, i.e.:
86 4 Stratification of Slurry Flow and Deposition of Solids in Pipes
dσ s
- = ρw gðSs- Sf ÞCvb ð4:2Þ
dz
D
σ s = ρw gðSs- Sf ÞCvb ðcos α - cos βÞ ð4:3Þ
2
D2
F N = ρw g ðSs- Sf Þ C vb ðsin β - β cos βÞ ð4:4Þ
2
D2
F W = ρw gðSs - Sf Þ Cvb ðβ - sin β cos βÞ ð4:5Þ
4
= ρw gðSs - Sf Þ C vb A2
FN 2ðsin β - β cos βÞ
= ð4:6Þ
FW ðβ - sin β cos βÞ
4.3 Modeling of Stratified Flows 87
The values calculated from Eq. (4.6) show that for β less than 60°, the total
normal force is only slightly greater than the submerged bed weight, while there is a
marked difference at larger values of β. With a deposit that completely fills the pipe,
β = π, and the normal force (FN) is twice FW.
A more rigorous analysis of the stresses (as in Matoušek and Krupička 2011)
gives a more general determination of the normal force (FN) that the bed exerts
against a pipe wall. It is derived from the local normal stress of solids at a pipe wall
(σ N) expressed from its vertical and horizontal components, which can differ in their
values. For the vertical solids stress (σ z), the assumption of the linear increase with
the depth below the top of the bed is taken (which agrees with Eq.(4.2)), and the
corresponding shear stress is negligible. The assumption of the linear relationship for
σ z(z) is justified by Shook and Roco (1991) by deriving Jansen’s equation for the
axial stress in the column of settled slurry in a vertical pipe as a function of depth.
The horizontal solids stress σ x = σ z/K, where K is a constant sensitive to granular bed
properties. If K = 1, then σ N = σ x = σ z, and Eqs. (4.3) and (4.4) hold. However, if
K < 1, then the normal stress at a pipe wall for principal stresses σ x and σ z is
determined using σ N = σ xsin2α + σ zcos2α as:
h i
D
σN = ρw gðSs- Sf ÞC vb ðcos α - cos βÞ Ksin 2 α þ cos 2 α ð4:7Þ
2
In the case of K = 1:
88 4 Stratification of Slurry Flow and Deposition of Solids in Pipes
D2 τ
F N = ρw gðSs- Sf ÞCvb ðsin β - β cos βÞ þ 12 0 Dβ ð4:10Þ
2 tan φ
The above equation is an extension of Eq. (4.4). For fully stratified flow (only a
few particles rolling and/or saltating over the top of the bed), the increase in the
solids stress due to colliding particles can be neglected. In a flow with a developed
shear layer, this portion of the solids stress may be significant.
The parameters used in an expression of the force balance for the layers are
schematically displayed in Fig. 4.5, and they are used in the momentum equations
for the upper layer and lower layer in Eqs. (4.11) and (4.12), respectively.
For the upper layer:
where τ1 is the fluid shear stress at boundary O1, τ12 is the shear stress at the
interface between the upper and the lower layer, and L is the distance over which Δp
is taken in the horizontal pipe. In the upper layer, both shear stresses oppose the
pressure gradient. For the lower layer:
where the total shear stress at the boundary O2 has two components for the fluid
and solids respectively: τ2 = τ2f + τ2s, where τ2s = μsFN/(O2L ).
In the force balance for the bed (the lower layer), the resisting force μsFN resulting
from the mechanical friction between the pipe wall and the particles comprising the
bed is in equilibrium with three other forces (Eq. (4.12)). One is the resisting force
due to fluid (viscous) friction of the bed interstitial liquid at the boundary O2 (i.e.,
τ2fO2L ). The remaining two are driving forces that tend to set the bed in motion. The
first of these occurs within the bed and is a downstream-directed force due to the
pressure differential (Δp = pA-pB in Fig. 4.5) acting on the portion of the cross-
section occupied by the bed, A2. This force is associated with the seepage
(or percolation) flow set up by the pressure gradient, Δp/L. The seepage flow creates
drag forces on the particles, which, in total, produce a force per unit length
4.3 Modeling of Stratified Flows 89
equivalent to Δp/L times the cross-sectional area occupied by the bed. The second is
the force resulting from the interfacial shear stress (τ12) imposed on the top surface of
the bed by the flow above it.
Unlike the solids shear stress (τ2s), the boundary fluid-like stresses are velocity
dependent and related to the flow velocity in each layer (V1, V2) through a boundary
friction coefficient ( f1, f2) (i.e., τ1 = f 1 ρf V 21 =8 and τ2f = f 2 ρf V 22 =8). The friction
coefficient f1 is solved by the law of the wall for carrying the liquid or homogeneous
suspension (e.g., formula by Churchill or Swamee-Jain as discussed in Chap. 2). The
case of f2 will be briefly outlined in Sect. 4.3.3. The interfacial shear stress (τ12) and
its friction coefficient ( f12) require more attention.
The top of the bed is an extraordinary boundary in the two-layer flow structure since
it moves if the bed slides. Furthermore, it is erodible. If the interfacial shear stress is
sufficiently high, the top of the bed gets eroded, and a shear layer develops. The
shear stress at the top of the bed is expressed in terms of the velocity
difference between the layers and the associated friction coefficient as f12 (i.e.,
τ12 = f12 ρf(V1 - V2)2/8).
Various formulae are in use for the interfacial friction factor ( f12). The common
assumption for these different formulae is that the bed surface is composed of coarse
particles and can be treated as a hydraulically rough boundary, as described by the
Nikuradse formula (Nikuradse 1933) adopted for the condition at the interface:
rffiffiffiffiffiffi
8 1 ε12
= ln ð4:13Þ
f 12 κ B Rhb
in which the friction coefficient depends on the ratio of the interface roughness
(ε12) to the hydraulic radius Rhb of the flow area above the bed associated with the
interface. The procedure described in Matoušek (2007) calculates the flow area and
corresponding Rhb and determines them from experimental data. κ and B are the
friction-law constants (κ = 0.4, B = 14.8).
If the top of the bed is virtually unsheared and is subject to virtually no erosion,
then it can be considered as a fixed boundary of the roughness ε12 related to the size
of the grains occupying the plane surface of the bed. This is usually the case if the
bed is composed of very coarse particles or the velocity of flow above the bed is very
low. In this case, the bed roughness (ε12) can be estimated to be approximately twice
the grain size (d ). If, however, the bed gets eroded and the shear layer develops of
several grain diameters in thickness, then the roughness increases considerably, and
another relation must apply to its determination. The increase in the roughness is
demonstrated in Fig. 4.6 where the interfacial friction coefficient ( f12) is related to
the delivered concentration (Cvd) for flow above a sand deposit (i.e., at velocities
90 4 Stratification of Slurry Flow and Deposition of Solids in Pipes
0.20
0.15
0.10
f12 (-)
0.05
0.00
0.001 0.010 0.100
Cvd (-)
Fig. 4.6 Interfacial friction coefficient versus delivered concentration for flow of slurry of 0.37 mm
sand above deposit in 150 mm pipe. Measured (points) and predicted (line) using Eq. (4.13) for ε12
twice grain size. (Matoušek & Krupička 2012)
below the deposition limit) in a pipe. At very low flow velocities, the Cvd is also very
low, showing that the erosion of the top of the bed is weak; its roughness is
successfully predicted by Eq. (4.13), with the roughness equal to double the size
of the sand particle.
If, however, the flow velocity increases, causing more erosion of the top of the
bed, and hence a higher Cvd, then an assumed value for roughness equal to twice the
particle size predicts considerably lower values of the friction coefficient than
obtained by the measurement. Apparently, the roughness must increase with Cvd
as well. The analysis of the roughness of the sheared bed shows that the roughness
ratio ε12/d is indeed not constant at high bed shear and increases primarily with the
dimensionless interfacial shear stress, called the bed Shields parameter, θ12 = τ12/
[(ρs - ρf)gd] (Wilson 1988, Wilson 1989, Wilson and Nnadi 1990). This relation
was further refined with an introduction of additional dimensionless numbers as the
population of experimental data expanded (e.g., Camenen et al. 2006, Matoušek and
Krupička 2009, 2014, Pugh and Wilson 1999).
It is worth noting that there are two configurations for which the interfacial friction is
not important, so that the other portions of the model can be verified without
requiring a value of f12.
One configuration is the fully stratified flow of a bed sliding down an inclinable
tube at a velocity very similar to the velocity of flow above the bed. Such flows were
4.3 Modeling of Stratified Flows 91
The models discussed below are generally applicable to turbulent flows of fully
stratified and partially stratified (heterogeneous) slurry flow composed of narrowly
graded solids in a Newtonian carrier at flow velocities above the deposition limit. Of
these, the Saskatchewan Research Council (SRC) model also covers broadly graded
solids, while the unified model can also be used for flow at velocities below the
deposition limit.
92 4 Stratification of Slurry Flow and Deposition of Solids in Pipes
Typically, these models predict the frictional pressure gradient, the position of the
interface, the velocities, and the concentrations in each layer for given Vm and Cvd of
the defined slurry in a pipe of known diameter and pipe roughness. Therefore,
besides their ability to predict a pipe characteristic curve, the models can provide
important information about the internal structure of the flow that is useful for other
purposes (e.g., for pipe wear prediction).
The Saskatchewan Research Council model adopted the principles of the early
model, and the SRC has been developing its components ever since (Gillies and
Shook 1994; 2000; Gillies et al. 1991, 2004; Shook and Roco 1991). The model has
a reputation for being reliable in a broad range of flow conditions, as its calibration
and validation model builds on a large experimental database collected for various
slurries in pipes of different sizes at SRC over more than half a century. The
modeling of boundary shear stresses incorporates Coulombic solids stress, Bagnold
solids stress (called kinematic dispersive stress in the model), and their reduction by
the off-the-wall hydrodynamic lift. In the two-layer structure, the upper layer
contains only suspended particles, and the lower layer is a combination of contact-
load and suspended-load particles. This is its main difference from the unified
layered model described below.
Recently, the model application was extended to broadly graded solids, making it
a multi-species model.
4.3 Modeling of Stratified Flows 93
A unified layered model (Matoušek and Krupička 2010, 2011) works as a one-layer
model if there is a deposit at the bottom of the pipe and as a two-layer model at
velocities above the deposition limit. Hence, it determines Vs and predicts flow
parameters at both sides of Vs with a smooth transition at Vs. This feature is enabled
by the fact that the same formulae are used for interfacial parameters in both regimes.
The model considers an interfacial transport layer embedded in the upper layer and
adjacent to the interface, in which particles are transported either as a contact load or
as a combined load (contact load and suspended load together). The bed, whether
stationary or sliding, contains exclusively contact-load particles. The distribution of
particles between the bed and the transport layer is determined by the interfacial
transport formula of the Meyer-Peter and Müller type.
Recently, the model was adopted to inclined stratified flows, and the results of its
predictions will be discussed in Chap. 7.
Other examples of existing layered models include three-layer models, such as those
by Doron and Barnea (1993) or Ramadan et al. (2005). Using three layers has a
particular advantage in inclined flows; in an upward-flowing line, conditions can
arise where the lowest layer slides backward down the pipe, while the two layers
above travel up the pipe.
Increasing the number of layers beyond three is not warranted, and a further
reduction in the granularity of the models is left to computational fluid dynamics
(CFD) and discrete element modeling (DEM) computations.
For turbulent flow, a non-Newtonian stratified model is essentially the same as those
described above, with the exception that the stresses associated with the boundary
conditions are now determined using methods suitable for the particular
non-Newtonian carrier fluid model being used, e.g., instead of using the Swamee-
94 4 Stratification of Slurry Flow and Deposition of Solids in Pipes
Jain equation to calculate pipe friction factor and hence, wall shear stress (τf), a
non-Newtonian variant suitable for the carrier fluid model must be used. See Darby
and Melson (1981), Metzner and Reed (1955), Swamee and Aggarwal (2011), or the
Wilson & Thomas method described in Sect. 6.2.2. The high viscosity and density of
the carrier means that turbulent suspension above the bed occurs more readily for
small particles, and since only low Reynolds number flow is obtained in industrial
pipelines, where the cross turbulence (v’) is small, large particles are rarely
suspended. The packing of the bed, flow under the bed, and flow at the interface
also vary somewhat from the Newtonian case, but these considerations are common
to laminar flow and so will be dealt with there.
In laminar flow, the transverse motions characteristic of turbulent flow do not exist,
so particles are free to settle into a bed, even when the carrier fluid exhibits a yield
stress (see Sect. 6.3.1). Acrivos’ group has reported resuspension due to particle/
particle collision (Leighton and Acrivos 1986, 1987; Schaflinger et al. 1990, 1995;
Zhang et al. 1992), which is restricted to regions of the pipe between the bed
interface and the dynamic center of the flow. Such behavior, however, is rarely
observed in hydraulic transport systems. Consequently, the flow of heterogeneous
suspensions under laminar conditions produces an almost perfect two-layer regime,
as shown in Fig. 6.13. For reasons explained in Sect. 6.3.1, it is possible to have a
plug, located around the dynamic center of the flow, that retains solids. However,
since these do not influence the flow at the boundaries, they do not influence the
various stresses at these boundaries, although they do increase the power consump-
tion through their mass.
Similar behavior is observed at the bed interface where, depending upon particle
type and local fluid viscosity, the solids may form a sheared layer a couple of particle
diameters deep and may be accounted for in a way similar to that described in Sect.
4.3.2. One difference, however, is the enhanced thickness of the viscous sublayer.
The effect is that many surfaces that would be rough in water-borne flows may be
considered to be hydraulically smooth. Where the particles are large, Graham et al.
(2014) have shown that their increase in τ12 can be modeled by reducing the
hydraulic diameter above the bed by the characteristic particle size.
The highly viscous nature of the fluid increases the size of the interstices between
particles through increased squeeze film and normal forces. As a result, the bed
concentration is lower than found in water-borne systems, both within the bed and
close to the pipe wall beneath the bed (i.e., along O2). The mechanical friction at the
wall is also modified by carrier rheology. At present, such changes are difficult to
quantify, but they can be obtained from the experiment. Fortunately, this is quite
simple, as in most stratified model formulations, the parameters Cvb and μs appear as
a lumped parameter, and pipe tests conducted for one condition can be used to
calibrate a more general model by manipulation of this lumped parameter. At the
time of writing, suitable bench-scale equipment for measuring this (without the need
to conduct pipe-loop tests) are under development at the CSIRO laboratories.
A more complete discussion of these flows is given in Sect. 6.3.
4.4 Prediction of Velocity at the Limit of Stationary Deposition 95
One of the major criteria which must be considered by pipeline designers and
operators in selecting the desirable operating range for a system concerns the
conditions under which a stationary bed of solids will occur in the line. In this
section, we focus on the prediction of this deposition or bedding condition. The
challenges presented by the operation of a slurry pipeline with a stationary deposit
will be discussed later in Chap. 5.
These conditions were first studied by Durand and his co-workers (Durand
1953; Durand and Condolios 1952; Gibert 1960), who defined the deposition
limit velocity (“vitesse limite de dépôt”) as the upper limit of mean velocity at
which a stationary deposit can be found.
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi Durand’s group defined a dimensionless
ratio F L = V s = 2 g ðSs =Sf - 1Þ D, in which Vs is the deposition limit velocity.
Durand and Condolios found the value of FL to be of the order of unity, and they
related it to particle size and volumetric concentration in a graph. Unfortunately,
two different versions of a graph were published in the 1950s, and they showed
rather different results. As discussed by Miedema (2016), the version in the
original research report and related article (Durand and Condolios 1952) predicts
lower values than the version for the international conference paper (Durand
1953). The more accessible conference version has been referred to frequently
in the literature, and by its comparison with more recent experimental results, this
1953 version developed a reputation for being strongly conservative in Vs
prediction. In many cases, beds begin to move at values considerably smaller
than predicted. However, the 1952 research report produces more reasonable
predictions. Note also that the graph (in either of the two versions) covers
volumetric concentrations up to 0.15 only; hence, neither graph can capture the
effect that a high concentration of solids actually has on Vs.
During the 1970s, work on the limiting velocity for stationary deposition shifted
from a data correlation approach to one based on a force balance on the bed, as
considered in a two-layer model. The prediction of Vs has since been further refined
with progress in the development of various layered models. In addition, a prediction
method for finer settling slurries, which do not exhibit distinct layers—even at the
velocity of the limit of stationary deposit—has been subject to investigation from the
late 1970s, starting with the work by Thomas (1979).
An updated version of the VSCALC algorithm for predicting the limiting velocity
for stationary deposition (originally provided in the second Edition of this book) is
presented in the form of a MATLAB script in Appendix A. It combines Wilson’s
approximations of force-balance results for medium and coarse particle sizes with
the Thomas formula for fine particles. These approaches are discussed in detail in the
following sections of this chapter.
96 4 Stratification of Slurry Flow and Deposition of Solids in Pipes
Alternative predictive methods have been developed, of which the SRC method (Shook
et al. 2013) relating FL with the Archimedes number, Ar = 4ρf gd3 ðρs - ρf Þ= 3μ2f ,
appears to be the most reliable. It is applicable to particles of Ar > 125, this
limiting value representing typical medium-sized sand in water.
4.4.2.1 Modeling
In principle, the point of incipient bed slip occurs when the forces tending to move
the deposit reach the point that they equal those resisting motion. If the driving forces
increase further, then the force balance is maintained and the bed travels en bloc as a
sliding bed if it does not break up. Thus, a force-balance analysis is required to
determine the limit of stationary deposition for stratified flows including heteroge-
neous flow. Any one of the existing layered models is a force-balance model and is
therefore suitable for the prediction of the deposition limit velocity in such flows. If a
layered model is not available, the pipeline designer can rely on other tools based on
the results of an early version of the two-layer model.
The results of Wilson’s early two-layer model for the limit of stationary deposi-
tion showed that the throughput velocity (Vm) at this limit is concentration depen-
dent, having small values at low concentration, which rises to a maximum (denoted
Vsm) at some intermediate concentration (depending on pipe size and particle size
and density) and then drops off again as the delivered solids concentration (Cvd)
approaches the loose-poured value for plug flow (Cvb). This behavior is shown
schematically in Fig. 4.7, in which the predicted deposition limit velocities (Vs)
form a locus delimiting the stationary deposit zone. The modeling also predicts the
velocities (Vmin) at which the pipe characteristic curve shows a minimum at each
delivered concentration (Cvd).
The conservative designer may be content to know only the maximum velocity at
the limit of deposition (Vsm), since maintaining the operating velocity above this
value ensures that deposition will not occur under any circumstances. In many cases,
the Vsm prediction is not too conservative, and this maximum value can generally be
recommended for design purposes. However, as discussed in Chap. 3, there are other
factors to consider before setting the minimum operating velocity for the transport
system.
Wilson’s two-layer model indicated that the predicted value of Vsm depends on
the number of variables, the most important being the internal pipe diameter (D), the
particle diameter (d ), and the relative solids density (Ss). These relationships were
expressed concisely in a nomographic chart reproduced here as Fig. 4.8 (Wilson
1979, Wilson and Judge 1978). This chart has gained considerable popularity and is
frequently used as a practical design aid. The nomograph’s characteristic arched
curve for particle size dependence was sometimes named by Wilson as the “demi-
4.4 Prediction of Velocity at the Limit of Stationary Deposition 97
Fig. 4.7 Definition sketch for the limit of stationary deposit zone in plot relating the dimensionless
pressure gradient (im) with the flow velocity (Vm). Boundary of shaded zone represents the
deposition locus for variable Cvd
Fig. 4.8 Nomographic chart for maximum velocity at limit of stationary deposition, from Wilson (1979)
Stratification of Slurry Flow and Deposition of Solids in Pipes
4.4 Prediction of Velocity at the Limit of Stationary Deposition 99
• The right-hand side of the curve indicates, conversely, that Vsm increases with
increasing particle size. This is typical for heterogeneous (partially-stratified)
flow where the different thicknesses of the bed for particles of different sizes
further complicate the determination of the force balance at the slip point of the
bed. Finer particles are more readily suspended, and thus, the bed is thinner than
is the case for coarser particles, where less particles are suspended. This affects
the resulting Vsm and causes the thinner beds of finer particles to stop sliding at a
lower velocity than thicker beds of coarser particles.
To demonstrate the particle size effect, consider particles of Ss = 2.65 in a pipe of
300 mm in diameter, as shown in Fig. 4.9. This diameter is located on the vertical
scale on the left-hand side of the chart and connected by a straight line to any desired
particle size on the arched curve. Vsm is then obtained by projecting this line to the
central vertical scale. For instance, a particle size of 5 mm (fully stratified flow) gives
Vsm of 2.6 m/s, and the same Vsm is found for a 0.2-mm particle on the other limb of
the particle-diameter curve (heterogeneous flow). For a 0.6-mm particle, the depo-
sition limit velocity increases to almost 4 m/s, which is the largest value found for
this pipe diameter and solids density.
A particularly useful feature of the nomographic presentation of results is that it
gives an immediate indication of the sensitivity of the output to variations in the
input. Thus, Fig. 4.9 shows that the value of Vsm for sand-weight solids in a 0.30 m
pipe is virtually unaffected by a variation of the particle diameter between 0.4 mm
and 1.0 mm. However, the right limb of the curve shows greatly increased sensitiv-
ity, so that a change in d from 0.15 mm to 0.20 mm alters Vsm by more than 25%.
100 4 Stratification of Slurry Flow and Deposition of Solids in Pipes
In Eq. (4.14), dmm is the particle size in millimeters, and the pipe size (D) is in
meters. Eq. (4.14) approximates the curve best for Sf = 1 and μs = 0.40 (Matoušek
1997).
As mentioned previously, the chart is based on an early version of the two-layer
model, which considered the bed roughness equal to a multiple of the particle size. It
appeared later that this model is not appropriate for stratified flows of medium-to-
coarse particles where the interfacial friction sets up a shear layer several grain
4.4 Prediction of Velocity at the Limit of Stationary Deposition 101
diameters in thickness at the top of the bed. Due to the presence of this shear layer,
the bed roughness is larger than the equivalent roughness of the unsheared bed
surface; thus, the resulting Vs must be smaller than the nomogram predicts for
particle sizes near the Murphian size (i.e., those that give the largest value of Vsm
for the pipe under consideration).
It is in accord with experimental evidence that for particles near the Murphian
size, the value of Vsm is less than that predicted by Fig. 4.8. Wilson (1992) suggested
to modify the chart results by introducing the dimensionless maximum deposition
limit velocity (Vsm,max). For conditions typical for sands, Vsm,max is calculated by a
simple power-law approximation:
0:13
V sm, max 0:018
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi = ð4:15Þ
2gDðSs =Sf - 1Þ ff
where ff is the friction coefficient for fluid alone. If the value of Vsm found from
the nomographic chart or Eq. (4.14) exceeds the value of Vsm,max from Eq. (4.15),
then the latter value should be used for Vsm.
4.4.2.2 Experiment
1
A simple electronic or thermal probe mounted in the bottom of the pipe, see Sect. 12.5.
102 4 Stratification of Slurry Flow and Deposition of Solids in Pipes
Experimental data from large pipes of industrial size are scarce, and the majority
of available data are from pipes of diameters up to 100 or 150 mm. Much more
experimental evidence is available for the effects of the particle size and the
delivered concentration on Vs than for the effect of the pipe size.
The thesis that there is a certain critical particle size at which Vs is maximal in any
particular pipe under consideration has been confirmed by a number of experiments,
recently by Kesely et al. (2017) for different fractions of glass beads (0.18 mm,
0.44 mm, 0.53 mm, and 0.9 mm) in a 100 mm pipe. This same work also looked at an
effect of the delivered concentration and confirmed that high delivered concentra-
tions of particles tend to decrease Vs in a pipe. The work compared the experimental
results with predictions of two Vs models discussed in Sect. 4.4.1: the VSCALC
model and two versions of the Durand model. A comparison of the VSCALC model
with the glass beads experiment exhibited a very good agreement except for Cvd
higher than 0.3, where the model tended to under-predict the deposition limit
velocity. Predictions using the 1952 version of the Durand graph agreed well with
the experimental results for Cvd up to 0.15, while the 1953 version considerably
over-predicted the deposition limit velocity at any concentration (Kesely et al. 2017).
Recent experiments conducted at the GIW Hydraulic Laboratory using a 203-mm
pipe enabled researchers to compare results for four different narrowly graded
fractions of solids of Ss = 2.65 transported in water (Matoušek et al. 2017).
Figure 4.10 shows the results expressed as the Durand parameter (FL) versus
Archimedes number (Ar). The figure reveals that the fraction of 0.8 mm sand
(Ar ≈ 2 × 104) indeed exhibits larger values of Vs than the two finer fractions
(0.025 mm and 0.12 mm) and one coarser fraction (7.5 mm). The scatter of
experimental data for FL of a certain fraction (approximately constant Ar) is due
partially to the effect of variable delivered concentration; note that Cvd is not
1.5
Experiment
Durand (1952)
SRC
1.0 VSCALC-V
FL (-)
0.5
0.0
10-1 101 103 105 107
Ar (-)
Fig. 4.10 Relationship between the dimensionless deposition limit velocity (FL) and Archimedes
number (Ar) for four individual fractions in 203 mm pipe. (Matoušek et al. 2017)
4.4 Prediction of Velocity at the Limit of Stationary Deposition 103
included in the formulation of FL or Ar. The plots also show a reasonable agreement
between the experiments and predictions by three different models.
The GIW experiments with these four solids fractions also included measure-
ments of the flow of slurries with fractions mixed together to form a broadly graded
slurry. The experiments showed that the deposition-limit velocity was not necessar-
ily lower in the flow of slurry composed of the four fractions than in the
corresponding flow of slurry of the one fraction with the highest deposition limit
velocity from the four (Matoušek et al. 2017). A disproportionate contribution of
individual fractions to the deposition limit velocity in broadly graded slurry flow was
taken into account in a multi-component model proposed in the same paper. The
predictive model took the VSCALC model as a basis for computation of the deposit
velocities of individual fractions.
This equation gives the minimum value of deposition limit velocity for small
particles (typically smaller than 70 μm), assuming turbulent flow. In Eq. (4.16), vf
represents the kinematic viscosity of the carrier fluid, and the coefficient of 9.0
applies in any consistent system of units.
For particles larger than the viscous sublayer and suspended by flow turbulence,
Wilson and Judge (1976) recommended the following equation:
V sm d
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi = 2:0 þ 0:3 log 10 ð4:17Þ
2gDðSs =Sf - 1Þ DC D,s
in which CD,s is the drag coefficient of a spherical particle of diameter d (see Sect.
2.5). Eq. (4.17) is limited to d/(DCD,s) > 10-5.
104 4 Stratification of Slurry Flow and Deposition of Solids in Pipes
Thomas (2015) modified Eq. (4.17) to improve the predictive ability of the
method in pipes of large diameters (d/(DCD,s) < 10-5),
V sm d
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi = 2:0 þ 0:305 log 10
2gDðSs =Sf - 1Þ DC D,s
ð4:18Þ
- 0:489 - 1:06
þ1:1 10 - 4 DCdD,s - 0:044 107 DCdD,s
Although the use of the deposit limit velocity (Vsm) ensures that operation with a
stationary bed will be avoided under any circumstances, occasionally, there are cases
where operating at throughput velocity less than Vsm may be both economically
attractive and deposit-free. Specifically, this occurs when the solids concentration at
the desired operating point is markedly different from that associated with Vsm, so
that the deposition limit velocity (Vs) at the desired concentration is significantly less
than Vsm. To deal with such cases, it is suggested to run a two-layer model for
stratified flow at the desired concentration and to predict the corresponding Vs.
Alternatively, Wilson’s technique of approximating the results of his two-layer
model can be used to obtain values of Vs for any delivered volumetric solids
concentration (Wilson 1986). The computational algorithm VSCALC incorporates
this technique, together with the nomographic chart and the associated equations
presented previously.
Another option is to use the generalized version of Wilson’s “demi-McDonald”
chart, as presented in Fig. 4.11. In this figure, the lines of constant concentration are
the result of averaging the model outputs, and they show a substantial reduction of Vs
with an increase in concentration. Furthermore, the Murphian particle size decreases
with increasing concentration. Larger particle sizes along the left sides of each curve
track along a similar line, but their positions still shift significantly downward.
For fine pseudo-homogeneous flows, Thomas’s formula (Eq. (4.16)) was
extended by Sanders et al. (2004) to include the effect of concentration (and particle
size). Thomas (2015) discussed the different equations developed for pseudo-
homogeneous flow and the criteria for the conditions of their applicability.
The stratified models used with non-Newtonian carrier fluids are essentially the same
as those used for Newtonian fluids, and they also exhibit a deposition locus, as
4.4 Prediction of Velocity at the Limit of Stationary Deposition 105
Fig. 4.11 Generalized nomographic chart for deposition limit velocity (Vs) at various delivered
concentrations. (Wilson et al. 2011)
described in Sect. 4.4.2 and Fig. 4.7. Figure 4.12 shows deposition loci for 1 mm
particles in a 300 mm pipe being conveyed by various non-Newtonian carrier fluids,
and where the coarse solids concentration ranges from 0 to the packed bed concen-
tration Cvb at plug flow. Unlike the models that use water, the shape and extent of
this deposition locus is sensitive to the carrier fluid rheology, especially in the case of
106 4 Stratification of Slurry Flow and Deposition of Solids in Pipes
3.0
D = 300mm, d = 1mm, Ss = 2.65, k = 3 Pa sn, n = 0.6
2.5 plug flow, Cv = Cb
Pressure gradient (kPa/m)
2.0
ty = 24 Pa
1.5 Vsm
44 34
52
1.0
0.5
Cvb = 0 (a)
0.0
0 0.2 0.4 0.6 0.8 1
Velocity, Vm (m/s)
9
D = 300mm, d = 1mm, Ss = 2.65, k = 3 Pa sn, n = 0.6
8
plug flow, Cv = Cb
7
5
ty = 24 Pa
Vsm
t0/ty (-)
4
34
3
44
2 52
1
Cvb = 0 (b)
0
0 0.2 0.4 0.6 0.8 1
Velocity, Vm (m/s)
Fig. 4.12 Effect of variation of the yield stress on the deposition locus and carrier fluid-only
friction loss curves. (a) expressed in terms of pressure gradient and (b) in terms of the ratio of carrier
fluid yield stress and wall shear stress, so that curves can be collapsed for clarity
3.0
D = 300mm, d = 1mm, Ss = 2.65, ty = 24 Pa, n = 0.6
2.5
Pressure gradient (kPa/m)
2.0
1.5 k = 2 Pa sn
4 3
5
1.0
0.5
0.0
0 0.5 1 1.5 2
Velocity, Vm (m/s)
Fig. 4.13 Effect of variation of the consistency index (k) on the deposition locus and carrier fluid-
only flow curve. (Full annotation is shown in Fig. 4.12b)
movement will also mean movement of the bed as well, providing a very robust
means of transporting coarse particles. However, this behavior does not come
without a cost, as increasing the yield stress also raises the carrier fluid-only curve
(as shown in Fig. 4.12a), and as a result, the suspension system curve too. It is also
worth noting in this figure the generally low-velocity values of the deposition locus.
It is generally possible to run the pumps up to speed and rapidly exceed these values
so that flow is usually assured without mishap. A similar story is told in Fig. 4.13
where the deposition locus retreats with an increase in k, at the expense of a higher
pressure drop overall.
In Fig. 4.14, if we make the fluid less shear-thinning by increasing n, the
deposition locus decreases—this time with a much smaller increase in pressure
gradient. However, n is generally a function of the material type, and so is not
readily modified (unlike τy and k, which can be easily manipulated by changing the
carrier concentration, pH, salt content, etc.).
While the parameter changes given above all modify the viscosity of the carrier
fluid, there is no simple way to combine each of these changes into a unified form.
For example, the use of a generalized Reynolds number based on the wall viscosity
(η0, given in Table 6.4) fails to provide any unified plots of deposition locus, and
hence Vs,max. Consequently, generalized modification to account for carrier fluid
rheology, as was done with the demi-McDonald for concentration in Sect. 4.4.4, is
not possible, and the designer must use stratified models based on their carrier fluids
rheology, as above, or seek other correlations.
Goosen (2014), studying the behavior of gold tailings, demonstrated that for
mildly non-Newtonian carrier fluids (τy < 2 Pa), deposition limit velocity
108 4 Stratification of Slurry Flow and Deposition of Solids in Pipes
3.0
D = 300mm, d = 1mm, Ss = 2.65, ty = 24 Pa, k = 3 kPa sn
2.5
Pressure gradient (kPa/m)
2.0
1.5 n = 0.5
0.7 0.6
1.0
0.9
0.5
0.0
0 0.5 1 1.5 2
Velocity, Vm (m/s)
Fig. 4.14 Effect of variation of the flow behavior index (n) on the deposition locus and carrier
fluid-only flow curve. (Full annotation is shown in Fig. 4.12b)
predictions developed for Newtonian fluids (Gillies et al. 2000, Sanders et al. 2004)
could be used for conservative estimates of the deposition limit velocity (typically
25% higher) for suspensions operating in the turbulent flow regime. Most cases
tested in the laminar flow regime experienced no deposition limit velocity of the
solids, although he found that, in some instances, stationary beds or sliding beds did
occur in agreement with the stratified modeling outlined above.
Evaluate the:
(a) Deposition limit velocity for transport of the tailings.
(b) Operational velocity for transport of the tailings.
(c) Effect of the presence of high-density zinc particles on solids deposition and
pipeline operation.
(a) Deposition Limit Velocity for Transport of Tailings
For safety reasons, it is appropriate to consider the worst-case deposition limit
velocity from the range of velocities for various possible concentrations of solids in
the slurry (i.e., Vsm). The diameter d85 is selected as an appropriate representative
particle size for the determination of Vsm. The size d85 is sufficiently below the
Murphian “worst-case” size to claim that Eq. (4.14), rather than Eq. (4.15), should be
used to calculate Vsm. If the sliding friction coefficient (μs) is estimated as 0.4 for
tailings particles at steel pipe wall, then:
of the same size but of lower density. The required operational velocity
Vm = 1.053.73 ≈ 3.9 m/s. If the line is operated at the velocity of 2.8 m/s proposed
for the tailings based on a solids density of 2650 kg/m3, it will develop a deposit of
considerable thickness that is composed of zinc particles leaked into the pipeline.
This may not be obvious at the pipeline outlet where tailings particles of all sizes,
including the coarsest particles, will continue to be discharged. However, the
reduction in the effective pipe diameter caused by the deposit will increase both
the hydraulic gradient and the required power if the flow rate is maintained constant.
If additional power is not available, the flow rate will be reduced.
To calculate the increase in the required pump head (Hw) resulting from the
required higher operating velocity, we can assume an estimated concentration of
transported solids Cv = 0.3. If the pseudo-homogeneous slurry is considered to be an
equivalent liquid (as in Case Study 3.1), then the required head Hw ≈ 11.9 m of
standard water (fitting losses neglected) to transport the tailings at 2.8 m/s. This
number is more than doubled to ensure deposit-free transport of tailings with the
addition of zinc particles with Hw ≈ 28.4 meters of standard water at 3.9 m/s.
Table 4.1 Deposition limit velocity Vsm for five candidate pipe sizes
D (in) 6 8 10 12 14
D (mm) steel schedule 40 s 154.1 202.7 254.5 304.8 336.6
Vsm (m/s) Eq. (4.15) 2.26 2.62 2.95 3.25 3.43
Vsm (m/s) Eq. (4.14) 2.80 3.31 3.79 4.21 4.46
Table 4.2 Vm (m/s) for Cvd (-)/D (mm) 154.1 202.7 254.5 304.8 336.6
Ms = 500 tonnes/hour
0.15 18.73 10.83 6.87 4.79 3.93
(bold text: selected
Vsm < Vm < 2Vsm) 0.20 14.05 8.12 5.15 3.59 2.94
0.25 11.24 6.50 4.12 2.87 2.36
0.30 9.37 5.41 3.43 2.39 1.96
0.35 8.03 4.64 2.94 2.05 1.68
solids concentrations appears economically viable and that these concentrations can,
in actual practice, be fed to the pump.
It is worth noting that a preliminary selection of appropriate pipe sizes can be
made, even if little is known about the particle size distribution of the transported
solids. Neither of the used equations employs a particle size (except for Eq. (4.14)
which is used for comparison only).
However, further analysis with additional criteria is necessary to finalize the pipe
size selection and complete the total system design. That analysis will be continued
in Case Study 5.1 Chap. 5.
References
B. Camenen, A. Bayram, M. Larson, Equivalent roughness height for plane bed under steady
flow. J. Hydr. Eng. 132(11), 1146–1158 (2006)
R.P. Chhabra, J.F. Richardson, Non-Newtonian Flow in the Process Industries (Butterworth-
Heinemann, Oxford, 1999)
R. Darby, J. Melson, How to predict the friction factor for flow of Bingham plastics. Chem.
Eng. J. 28, 59–61 (1981)
P. Doron, D. Barnea, A three-layer model for solid-liquid flow in horizontal pipes. Int. J. Multiph.
Flow. 19, 1029–1043 (1993)
Durand R, Basic relationships of the transportation of solids in pipes – Experimental research.
Presented at the international association for hydro-environment engineering and research
congress, Minneapolis, 1953
R. Durand, E. Condolios, Donnees techniques sur le refoulement hydraulique des matériaux solides
en conduite. Journées de l’Hydraulique 2, 27–55 (1952)
R. Gibert, Transport hydraulique et réfoulement des mixtures en eonduites. Annales des Ponts et
Chausées 130(4), 437–494 (1960)
R.G. Gillies, C.A. Shook, Concentration distributions of sand slurries in horizontal pipe flow. Part.
Sci. Tech. 12, 45–69 (1994)
R.G. Gillies, C.A. Shook, Modelling high concentration settling slurry flows. Can. J. Chem. Eng.
33, 709–716 (2000)
R.G. Gillies, C.A. Shook, K.C. Wilson, An improved two layer model for horizontal slurry pipeline
flow. Can. J. Chem. Eng. 69, 173–178 (1991)
R.G. Gillies, J. Schaan, R.J. Sumner, M.J. McKibben, C.A. Shook, Deposition velocities for
Newtonian slurries in turbulent flow. Can. J. Chem. Eng. 78, 704–708 (2000)
R.G. Gillies, C.A. Shook, J. Xu, Modelling heterogeneous slurry flows at high velocities.
Can. J. Chem. Eng. 82, 1060–1065 (2004)
P. Goosen, Trends in stationary deposition velocity with varying slurry concentration covering the
turbulent and laminar flow regimes. Presented at the 19th international conference on
hydrotransport, Golden, 2014
L.J.W. Graham, L. Pulllum, J. Wu, Flow of non-Newtonian fluids in pipes with large roughness.
Presented at the 19th international conference on hydrotransport, Golden, 2014
M. Kesely, V. Matoušek, P. Vlasák, Coarse particles conveying in horizontal and inclined pipes.
Presented at the 5th international symposium on reliable flow of particulate solids, Skien, 2017
D. Leighton, A. Acrivos, Viscous resuspension. Chem. Eng. Sci. 41, 1377–1384 (1986)
D. Leighton, A. Acrivos, The shear-induced migration of particles in concentrated
suspensions. J. Fluid Mech. 181, 415–439 (1987)
V. Matoušek, Flow Mechanism of Sand-Water Mixtures in Pipelines (Delft University Press, Delft,
1997)
References 113
V. Matoušek, Interaction of slurry pipe flow with a stationary bed. J. South Afr. Inst. Min. Metall.
107, 365–372 (2007)
V. Matoušek, J. Krupička, On equivalent roughness of mobile bed at high shear stress. J. Hydrol.
Hydromech. 57, 191–199 (2009)
V. Matoušek, J. Krupička, Modeling of settling-slurry flow around deposition-limit velocity.
Presented at the 18th international conference on hydrotransport, Rio de Janeiro, 2010
V. Matoušek, J. Krupička, Unified model for coarse-slurry flow with stationary and sliding bed.
Presented at the 15th international conference on transport and sedimentation of solid particles,
Wroclaw, 2011
V. Matoušek, J. Krupička, Intense bed-load friction and transport: The effect of concentration.
Presented at river flow 2012: 6th International conference on fluvial hydraulics, San Jose, 2012
V. Matoušek, J. Krupička, Analysis of concentration profiles in dense settling-slurry flows.
Presented at the ASME fluids engineering division summer meeting FEDSM, Incline
Viallage, 2013
V. Matoušek, J. Krupička, Interfacial friction and transport in stratified flows. in Proceedings of the
Institution of Civil Engineers-Maritime Engineering, vol 167 no 3 (2014), pp. 125–134
V. Matoušek, R. Visintainer, J. Furlan, G. McCall, A. Sellgren, Deposition limit velocity: Effect of
particle size distribution. Presented at the 18th International conference on transport and
sedimentation of solid particles, Prague, 2017
A.B. Metzner, A.R. Reed, Flow of non-Newtonian fluids-correlation of the laminar, transition, and
turbulent-flow regions. AICHE J. 3, 434–440 (1955)
S.A. Miedema, Slurry Transport: Fundamentals, A Historical Overview & The Delft Head Loss &
Limit Deposit Velocity Framework (SAM-Consult, Delft, 2016)
J Nikuradse, in Strömungsgesetze in rauhen rohren (VDI-Forschungsheft, 1933). p. 361
F.J. Pugh, K.C. Wilson, Role of the interface in stratified slurry flow. Powder. Tech. 104, 221–226
(1999)
A. Ramadan, P. Skalle, A. Saasen, Application of a three-layer modeling approach for solids
transport in horizontal and inclined channels. Chem. Eng. Sci. 60, 2557–2570 (2005)
R.S. Sanders, R. Sun, R.G. Gillies, M.J. McKibbem, C. Litzenberger, C. Shook, Deposition
velocities for particles of intermediate size in turbulent flow. Presented at the 16th International
conference on the hydraulic transport of solids, Cranfield, 2004
U. Schaflinger, A. Acrivos, K. Zhang, Viscous resuspension of a sediment within a laminar and
stratified flow. Int. J. Multiphase Flow 16, 567–578 (1990)
U. Schaflinger, A. Acrivos, H. Stibi, An experimental study of viscous resuspension in a pressure-
driven plane channel flow. Int. J. Multiph. Flow 21, 693–704 (1995)
C.A. Shook, M.C. Roco, Slurry Flow: Principles and Practice (Butterworth-Heinemann, Oxford,
1991)
C.A. Shook, R.G. Gillies, R.S. Sanders, R. Spelay, Slurry Pipeline Systems – Short Course Notes
(Saskatchewan Research Council, SRC Publications, Saskatoon, 2013)
P.K. Swamee, N. Aggarwal, Explicit equations for laminar flow of Bingham plastic fluids. J. Pet.
Sci. Eng. 76, 178–184 (2011)
A.D. Thomas, Predicting the deposit velocity for horizontal turbulent pipe flow of slurries.
Int. J. Multiph. Flow 5, 113–129 (1979)
A.D. Thomas, A modification of the Wilson & Judge deposit velocity equations, extendings its
application to fine particles and larger pipe sizes. Presented at the 17th international conference
on transport and sedimentation of solid particles, Delft, 2015
E.J. van Riet, V. Matoušek, S.A. Miedema, Theoretical description and numerical sensitivity
analysis of Wilson model for hydraulic transport of solids in pipelines. J. Hydrol. Hydromech.
44, 208–222 (1996)
K.C. Wilson, Slip point of beds in solid-liquid pipeline flow. J. Hydr. Div. ASCE 96, 1–12 (1970)
K.C. Wilson, A unified physically-based analysis of solid-liquid pipeline flow. Presented at the 4th
International conference on the hydraulic transport of solids in pipes, Banff, 1976
114 4 Stratification of Slurry Flow and Deposition of Solids in Pipes
K.C. Wilson, Deposition-limit nomograms for particles of various densities in pipeline flow.
Presented at the 6th International conference on the hydraulic transport of solids in pipes,
Canterbury, 1979
K.C. Wilson, Effect of solids concentration on deposit velocity. J. Pipelines 5, 251–257 (1986)
K.C. Wilson, Analysis of bed-load motion at high shear rates. J. Hydr. Eng. ASCE 113, 97–103
(1987)
K.C. Wilson, Evaluation of interfacial friction for pipeline transport models. Presented at the 11th
International conference on the hydraulic transport of solids in pipes, Cranfield, 1988
K.C. Wilson, Mobile-bed friction at high shear-stress. J. Hydr. Eng. ASCE 115, 825–830 (1989)
K.C. Wilson, Influence of particle properties on solids effect. in Proceeding of the 10th Interna-
tional Kol. Massenguttransport durch Rohrleitungen, Universität Paderborn (Meschede, 1992)
K.C. Wilson, N.P. Brown, Analysis of fluid friction in dense-phase pipeline flow. Can. J. Chem.
Eng. 60, 83–86 (1982)
K.C. Wilson, D.G. Judge New techniques for the scale-up of pilot plant results to coal slurry
pipelines. in Proceeding of the International Symposium on Freight Pipelines, University of
Pennsylvania (Philadelphia, 1976). pp. 1–29
K.C. Wilson, D.G. Judge, Analytically-based nomographic charts for sand-water flow. Presented at
the 5th international conference on the hydraulic transport of solids in pipes, Hannover, 1978
K.C. Wilson, F.N. Nnadi, Behaviour of mobile beds at high shear stress. in Proceeding of the 22nd
International Conference on Coastal Engineering, vol. 3, (Delft, 1990) pp. 2536–2541
K.C. Wilson, M. Streat, R. A. Bantin, A formula for the velocity required to initiate particle
suspension in pipeline flow. Presented at the 2nd international conference on the hydraulic
transport of solids in pipes, Coventry, 1972
K.C. Wilson, G. Addie, A. Sellgren, R. Visintainer, Simplified approach to effect of concentration
on deposition limit. in Presented at the 15th International Conference on Transport and
Sedimentation of Solid Particles (Wroclaw, 2011)
K. Zhang, A. Acrivos, U. Schaflinger, Stability in a two-dimensional Hagen-Poiseuille
resuspension flow. Int. J. Multiph. Flow 18, 51–63 (1992)
Chapter 5
Settling Slurry Flow
5.1 Introduction
The three basic types of settling slurry flow were described in Chap. 3. At one
extreme (the anterior end of the elephant, as it were) are large, rapidly settling
particles. The submerged weight of these particles cannot be carried by fluid support
mechanisms and must be transferred downwards by continuous or sporadic inter-
granular contacts. When all particles behave in this way, the flow is fully stratified.
The other limiting case is represented by the behavior found for fine sand slurries,
known as pseudo-homogeneous flow. In this type of flow, the particles are carried by
the fluid rather than by inter-granular contacts. The intermediate case is known as
heterogeneous flow, for which both intergranular contact and fluid support mecha-
nisms are significant. A special case is the flow of broadly graded solids in a
Newtonian carrier in which grains supported by different mechanisms interact with
each other. Another, although rather unusual, case is slurry flow above stationary
deposit at the bottom of a horizontal pipe where the top of the deposit poses an
additional boundary for the flow.
A Historical Perspective of Settling Slurry Flow Models
From the early years of investigations into the hydraulic transport of solids in
pipelines, different types of slurry flows were believed to require different forms
of predictive formulae to calculate slurry flow friction. Naturally, the early formulae
were of empirical origin. The first predictive correlations for different types of
settling slurry flows were formulated in the 1950s, with the major activities being
conducted by a number of research groups, primarily in France, the United King-
dom, and the Soviet Union.
For pseudo-homogeneous flows, the equivalent liquid concept was accepted as a
standard (Durand 1951; Newitt et al. 1955), and it is still found useful today. A
model based on this concept will be scrutinized in Sect. 5.2.
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 115
R. Visintainer et al., Slurry Transport Using Centrifugal Pumps,
https://doi.org/10.1007/978-3-031-25440-6_5
116 5 Settling Slurry Flow
Two research groups made major progress in formulating functional relations for
flows of slurries coarser than those in pseudo-homogeneous flow. The French group
(from the research center Sogreah), with R. Durand as its most frequently published
author (Durand 1953; Durand and Condolios 1952, and other works later summa-
rized by Gibert 1960) collected an extensive experimental database containing data
for slurry flow of quite different solids fractions (from fine sand to coarse gravel,
0.18 to 22.5 mm) and pipe sizes (from 40 to 580 mm). In their semi-empirical
processing of the collected data, the group expressed the excess gradient in hetero-
geneous flow by means of the dimensionless ratio (Φ) defined by Φ = (im - if)/
(Cvd if). The number Φ is constant for pseudo-homogeneous flows according to the
equivalent liquid model (ELM), but it varies with particular conditions in coarser
flows. To express this variation, Φ was related to another dimensionless variable (Ψ )
which included the flow- and particle-related quantities expressed as a product of
two Froude numbers. Curve fitting of the group’s experimental data resulted in
Φ = KΨ n, with n = -1.5. Different expressions for Ψ were used in the group’s
publications, with the most complete form of the Φ - Ψ correlation being:
pffiffiffiffiffin
im - if V 2m gd
=K ð5:1Þ
C vd if gDðSs =Sf - 1Þ vt
For Eq. (5.1), the experimental data collected by the Sogreah group are well fitted
with K = 82 (where n = -1.5). The correlation (in various modifications) has gained
considerable popularity as a design tool because of its usefulness and simplicity, and
it is still used for industrial purposes (van den Berg 2013). Following the establish-
ment of this correlation, a growing corpus of experimental data has provided
evidence that although Eq. (5.1) is not theoretically sound, it may still be useful,
particularly for heterogeneous flows in the approximate range of 4 < Ψ < 15
(Matoušek 1997, van den Berg 2013). The heterogeneous flows are discussed in
Sect. 5.4.
At Imperial College in London, Newitt’s group utilized the concept of mechanical
sliding friction between moving solids and the pipe to recognize the difference
between heterogeneous and fully stratified flow, not directly considered in Durand’s
correlation, Eq. (5.1). Based on a force-balance analysis of the stratified flow, Newitt
et al. (1955) found that the Durand equation must take n = -1.0 instead of -1.5 to
satisfy the fully stratified flow conditions. The force-balance considerations also
justified a use of another different form of the excess pressure gradient than the one
used by Durand et al. This version of the dimensionless number reads (im - if)/
[Cvi(Ss - Sf)], and for practical predictions, it is modified by replacing Cvi with Cvd
to obtain (im - if)/[Cvd(Ss - Sf)]. Newitt et al. suggested that this ratio (called the
relative solids effect) was constant (equal to 0.8) for fully stratified flow. pffiffiffiffiffi
A similar result would be obtained by Eq. (5.1) if n was set to -1.0 and vt = gd
reached a virtually constant value independent of the grain size, as is the case for
large, heavy particles like gravel grains where the drag coefficient CD has a constant
value (see Fig. 2.21). Unfortunately, experiments by Newitt’s group were limited to
5.2 Pseudo-homogeneous Flow 117
0.4
im (-)
0.3
0.2
0.1
0
0 2 4 6 8
Vm (m/s)
Fig. 5.1 Pipe flow curves for heterogeneous flow (black) and stratified flow (red) of granite rock
slurry in 203-mm pipe. (Experimental data from GIW Hydraulic Laboratory)
one small pipe with an inner diameter of two inches (approximately 50 mm); hence,
the group’s pioneering theoretical findings could not be sufficiently verified by their
own experiments. The fully stratified flows are discussed further in Sect. 5.3.
Many authors compared their data sets with Durand’s and Newitt’s correlations
and have shed more light on their range of application (see, e.g., the review by
Miedema (2016)). Figure 5.1 compares typical experimental results for the friction
loss in heterogeneous slurry flow with that in coarse-stratified slurry flow (both
slurries narrowly graded) from a laboratory loop, with predictions by Eq. (5.1). It
demonstrates that the Durand friction loss model can be successful in predicting
narrowly graded heterogeneous slurry flows, while it requires modifications if
applied to strongly stratified coarse slurry flows.
Further details of the historical perspective can be found in the earlier editions of
this book.
Sm
im = i and jm = jf ð5:2Þ
Sf f
in which the specific gravity of the slurry (Sm) is typically related to the delivered
concentration of solids (Cvd). In pseudo-homogeneous flows, the difference between
Cvi and Cvd can be neglected at flow velocities well above the deposition limit;
hence, the concentration to be used is not an issue. However, another aspect, related
to the use of the slurry density in the ELM, requires more attention. Some experi-
ments show that for some pseudo-homogeneous flows, the dimensionless pressure
gradient of the slurry (im) exceeds that of the liquid (if) by a smaller margin than
given by Eq. 5.2. In order to empirically model the effect of this solids friction
reduction, a generalized version of ELM employs the reducing parameter A′:
In this equation, A′ values are between 0 (negligible solids friction) and 1 (resulting
in Eq. (5.2)) and are subject to calibration by experimental data for scale-up
purposes. Expressed in terms of the relative solids effect on the total frictional
head loss, the generalized form of ELM reads:
im - if im - if i
= = A0 f ð5:4Þ
Sm - Sf C vd ðSs - Sf Þ Sf
1.0
(im-if)/[Cvd(Ss-Sf)]
0.1
Cv12.5%, Sand 1
C=
Cv25%, Sand 1
C=
Cv34%,Sand 1
C=
Cv17%, Sand 2
C=
Cv17%, Sand 2
C=
Equivalent fluid model, D =D0.10 m
0.0
1 2 3 4 5 6 7 8 9 10
Vm (m/s)
Fig. 5.3 Relative solids effect for 85 μm and 400 μm sands in a 100-mm pipe, after Whitlock et al.
(2004)
Another experiment (Whitlock et al. 2004), this one with a different fine sand
fraction in a different pipe—85 μm grains in a pipe loop with an internal diameter of
100 mm—showed a similar result; see Fig. 5.3. In the plot, the line represents the
equivalent liquid model (Eq. (5.4) with A′ = 1). Note that data for the 85 μm sand
(Sand 1) closely follows the equivalent liquid line for velocities above 1.5 m/s.
120 5 Settling Slurry Flow
Figure 5.3 also examines experimental data for a medium 400 μm sand in the
same pipe (Sand 2). Its behavior is quite different, exhibiting a relative solids effect
that decreases with the increasing throughput velocity (Vm), following a straight line
on the logarithmic plot of the figure. This represents a heterogeneous flow behavior.
Before describing this behavior in detail in Sect. 5.4, we must first examine the
extreme case opposite of pseudo-homogeneous flow: the fully stratified flow with a
sliding bed.
For engineering purposes, the most complete, sophisticated, and perhaps accurate
evaluation of fully stratified flow should be based on solving the layered model
outlined in Chap. 4. If such an evaluation is not possible, then alternative simplified
methods can be employed, as described in the following sections.
Let us first consider a limiting case of coarse flow, with the concentration
approaching that of loose-packed particles (Cvb). Such a flow is known as plug
flow or coarse-particle dense-phase flow. In this case, the subtended half-angle (β) to
the top of the bed (lower layer) approaches π, giving FN (in the configuration of the
original Wilson two-layer model) equal to twice the submerged weight of the
particles. The force balance for these conditions predicts that the dimensionless
pressure gradient required to set the plugin motion (ipg) is given by
This velocity effect arises from the variation of the lag ratio (Λ), which is related
to the ratio of in situ to delivered concentration (Eq. 3.5) and depends chiefly on the
ratio (Vr) between the throughput mixture velocity and the velocity at the limit of
deposition (i.e., Vr = Vm/Vsm). Thus, the limit of deposition is important not only in
its own right but also as a parameter that influences the solids effect at larger
velocities within the normal operating range.
To simplify the use of the original two-layer model for fully stratified flow and to
provide an easily used tool for first estimates in a pipeline system design, the chart,
shown below, was created (Fig. 5.4). This chart uses typical values of input
quantities to the model.
For any given delivered solids concentration, the difference between the mixture
pressure gradient and that for an equal flow of fluid alone depends not only on the
relative concentration (Cr) but also on the velocity ratio (Vr). In Fig. 5.4, the excess
pressure gradient has been expressed relative to the dense particulate plug gradient,
as given by Eq. (5.5). This relative excess pressure gradient, denoted ζ, is defined as
ζ = (im - if)/ipg. Thus, Fig. 5.4 plots ζ versus Vr, with Cr as a parameter. At large
values of Vr, the curves flatten, approaching asymptotic values that represent the
relative excess gradient evaluated at velocities high enough to eliminate any hold-up
of solids.
Cvd = 0.15
Cvd = 0.20
Cvd = 0.29
Cvd = 0.30
0.1
1 10
Vm (m/s)
Fig. 5.5 Relative solids effect for 6.1 mm granite rock in a 203-mm pipe. Open symbols used for
Vm ≤ Vsm. (Experimental data from GIW Hydraulic Laboratory)
We refer to this as the Vsm model for friction loss. In the original Wilson-Addie
formula and in Fig. 5.4, it was assumed that μs = 0.40.
A force-balance-based analysis of fully stratified flow justifying the form of
Eq. (5.6) is outlined in Matoušek et al. (2015). It explains the use of the value of
0.25 for the exponent of the velocity ratio, which is associated with an expression for
the friction coefficient. The appropriateness of the 0.25 exponent for fully stratified
flows has also been confirmed by laboratory experiments (e.g., Matoušek 1997,
Matoušek et al. 2015).
Recent experiments at the GIW Hydraulic Laboratory in 103-, 203-, and 489-mm
pipes with very coarse slurries also supported the value of 0.25 (Visintainer et al.
2017, 2022). The experiments carried out in 103 and 203 mm pipes used the same
grade of crushed granite rock at various concentrations (Cvd). Figure 5.5 compares
the relative solids effect for slurries of four different concentrations of the rock in the
203-mm pipe and demonstrates a lack of sensitivity to Cvd in the entire range of flow
velocities (Vm). The trend in the relationship between the relative solids effect and
Vm is well approximated by the slope 0.25.
Figure 5.6 plots the relative solids effect versus the relative velocity (Vm/Vsm) for
the same slurry in the pipes of the two different sizes. The introduction of the
reference velocity (Vsm) is required in order to generalize the relationship between
the relative solids effect and the flow velocity. The relative velocity enables all data
5.3 Fully Stratified Flow with Sliding Bed 123
0.25
1
0.1
0.5 5
Vm/Vsm (-)
Fig. 5.6 Relative solids effect for 6.1-mm granite rock in 203 mm and 103 mm pipes. Open
symbols used for Vm ≤ Vsm. Experimental data from GIW Hydraulic Laboratory
0.1
0.1 1 10
Vm /Vsm (-)
to collapse to one band in the plot, while two separate bands of data points would
result if the solids effect was plotted against Vm instead of Vm/Vsm on the horizontal
axis.
Figure 5.7 expands the plot with additional experimental data from the GIW
experiments in the larger pipe of 489 mm inner diameter using the same abscissa as
Fig. 5.6. Again, the data form one band, exhibiting a tight correlation between the
two dimensionless groups.
The GIW experiments confirmed that the deposition limit velocity (Vsm) is a
suitable reference velocity for very coarse-stratified slurry flows. Moreover, the
experiments validated the Vsm model (composed of Eq. (5.6) and the corresponding
equation for Vsm), which predicted the line in Fig. 5.7 taking μs = 0.48 for a sliding
bed of crushed granite rock in a steel pipe. Note that the model is considered for
deposit-free flows only (i.e., for flows at Vm > Vsm).
124 5 Settling Slurry Flow
Fig. 5.8 Relative solids effect for 0.88 mm and 6.1 mm granite rocks in 203 mm pipe. Open
symbols used for Vm ≤ Vsm. Experimental data from GIW Hydraulic Laboratory
Figure 5.8 plots the results for the coarse granite rock together with the results for
a finer (heterogeneous flow) fraction in the same 203-mm pipe loop. The heteroge-
neous flow seems to exhibit a power-law relation with Vm, but the value of the power
is considerably higher than 0.25.
Fully stratified flow exhibits a higher energy consumption than other settling
slurry flow types. Its relative solids effect, which is dominated by contact load, can
be thought of as representing the upper limit of excess pressure gradient. As the
experimental data in Fig. 5.8 suggest, heterogeneous slurry flows will generally
display smaller values of the solids effect. In the following section, we will inves-
tigate the fluid suspension of particles and show how this suspended load combines
with contact load in order to model heterogeneous slurry flow.
s
2
M
1
log
Vsm V0 V50 V100 Vm
Fig. 5.9 Schematic log-log plot of relative solids effect versus flow velocity
We refer to this as the V50 model for friction loss. Note that the concentration ratio
(Cc/Cvd), called the stratification ratio, can be estimated as Cc/Cvd = 0.5(V50/Vm)M
from Eq. (5.7).
Essentially, Eq. (5.7) contains two parameters, V50 and M, which can be evaluated
directly by pipe-loop experimentation. As mentioned previously, measured values of
the solids effect and corresponding measured Vm can be plotted on a log-log chart
similar to that of Fig. 5.9 and approximated to a straight line. The slope of the line
gives M, and the value of Vm corresponding to a solids effect of μs/2 gives V50. This
is the preferred way to determine the values of these parameters for the heteroge-
neous flow of a particular slurry. If the experimental values are not available, then
values based on previous experimental evidence should be used. The power M is
5.4 Heterogeneous Flow 127
typically close to 1.0 for slurries with a narrow particle grading (e.g., see the slope
for the heterogeneous flow data in Fig. 5.8).
Investigations into the factors that influence V50 were discussed in detail in earlier
editions of this book and resulted in rather complicated equations for determining
V50. A simpler, and relatively successful, approximation proposed for V50 ignores
minor variations of ff and d/D and uses a power-law equation with d50 (Ss – Sf) and
the relative viscosity (νr). The resulting evaluation formula reads
d 0:35
Ss - Sf
V 50 = 44:1 50
ð5:8Þ
ν0:25
r 1:65
For a slurry composed of typical sand solids in water of standard density (Ss – Sf)
equals 1.65, and the bracketed portion of Eq. (5.8) equals 1.00. The relative viscosity
(νr) represents the ratio of the actual viscosity of the carrier fluid to that of water at
20 °C. The value of V50 obtained from Eq. (5.8) is substituted into Eq. (5.7) to obtain
the solids effect (im – if), where if is the gradient for an equal flow of carrier fluid.
1
D = 103mm, d50 = 0.71mm, d85 = 2.55mm, Cvd = 0.19
(im - if)/Cvd (Ss - Sf) (-)
0.1
V50 0
Vm (m/s)
Fig. 5.10 Relative solids effect for medium-sized granite rock. Points show experimental data.
Line produced from Eqs. (5.7) and (5.8) with μs = 0.48 and M = 1.0. Open symbols used for
Vm ≤ Vsm. (Experimental data from GIW Hydraulic Laboratory)
viewed as the sum of the fluid component (if), and the solids effect (im – if). The fluid
component scales inversely with the diameter ratio, but the solids effect is indepen-
dent of pipe diameter, and thus forms a further parameter to be kept constant in the
scale-up process. Velocity scaling does not appear to be greatly influenced by the
solids effect and should be based on the friction factor for the fluid component only.
Figure 5.10 compares, on logarithmic coordinates, experimental data obtained at
the GIW Hydraulic Laboratory for an aqueous slurry of 0.7-mm granite rock in a
0.103-m pipe with predictions for the same slurry using Eq. (5.7) where M = 1 and
V50 is determined from Eq. (5.8). The plotted points fall on an essentially straight
line and are quite reasonably matched by the model prediction for Vm > Vsm. The
experimentally determined value of V50 = 3.64 m/s and is determined by finding the
Vm at which (im - if)/[Cvd(Ss - Sf)] = μs/2 = 0.24. This is very close to the 3.78 m/s
predicted by Eq. (5.8).
With V50 and M determined experimentally for the 0.103-m pipe, scale-up to a
larger pipe diameter can be carried out. The larger diameter of 0.203 m has been
selected because experimental data are available with the same solids fraction in this
larger pipe; thus, the scaled-up results can be verified directly (although the sample
used showed slightly different particle size characteristics, specifically
d50 = 0.88 mm, and d85 = 2.63 mm).
The experimental data confirm that V50 and M are virtually the same for either
pipe because both sets are part of the same population in the plot of Fig. 5.11. This
result is consistent with Eq. (5.8) for V50, which does not contain D as a variable, and
with the suggestion of constant M being unaffected by pipe size.
5.4 Heterogeneous Flow 129
Fig. 5.11 Relative solids effect for medium-sized granite rock in 103 mm and 203 mm pipes. Open
symbols used for Vm ≤ Vsm. (Experimental data from GIW Hydraulic Laboratory)
D = 103 mm
D
D = 203 mm
D
(im-if)/Cvd(Ss-Sf) (-)
1 D = 489 mm
0.1
0.1 1
Vm /V50 (-)
Fig. 5.12 Relative solids effect for heterogeneous flows in three pipe sizes. Line produced from
Eqs. (5.7) and (5.8) with μs = 0.48 and M = 1.0. Open symbols used where Vm ≤ Vsm. (After
Matoušek et al. 2020)
M
f f 2 μs V
i m = Sf V þ C ðS - S Þ 50 ð5:9Þ
2gD m 2 vd s f V m
The velocity at the minimum points can be obtained by differentiating Eq. (5.9)
with respect to Vm and setting the result equal to zero. For the case dealt with here,
where it is assumed that ff and V50 are independent of pipe diameter, and the effect of
Vm on ff is neglected, it can be readily shown that the value of Vm at the minimum is
Mþ2
1
C vd ðSs - Sf ÞgDμs V M
50 M
V min = ð5:10Þ
2Sf f f
For a given pipe size and M = 1.0, this velocity depends on Cvd0.33, and for a
given delivered concentration, it scales with pipe size according to D0.33. A numer-
ical example involving the minimum point of the system curve is given in Case
Study 5.1.
The 4-Component Model (4CM) was developed for use with settling slurries of any
grading profile, from broad to narrow to bimodal. The model assumes that the fine
(typically <40 μm) particles combine with the liquid phase to form a carrier fluid
having Newtonian characteristics and that the total mixture flow is both turbulent
and at a velocity above the limit of stationary deposition.
The solids are divided into size-delimited fractions (or components), and the
incremental contribution of each is evaluated according to one of the previously
5.5 Broadly Graded Settling Slurry Flow 131
Fig. 5.13 Solids fractions for model components delineated along the particle size distribution
curve
defined models. These contributions are then summed to predict the total dimen-
sionless pressure gradient in the pipe:
where the subscripts f, p, h, and s refer to the carrier fluid (carrier liquid plus fine
solid fraction), pseudo-homogeneous, heterogeneous, and fully stratified compo-
nents, respectively.
The volume of each solids component is defined by a fraction (X) of the total
solids volume in the slurry, such that the combined fractions sum to unity:
Xf þ Xp þ Xh þ Xs = 1 ð5:12Þ
has been applied in the following treatment; however, a more generalized formula-
tion could also be developed.
In the development of the 4CM, individual components are not treated in
isolation. Rather, each solids component in each sub-model is evaluated as being
carried or conveyed by a mixture of the liquid plus the finer solids components,
which act as its carrier fluid. For example, the fully stratified component solids
represented by Xs are considered as being conveyed by a mixture of the liquid and
the three finer solids components Xh, Xp, and Xf. In a similar fashion, this three-
component mixture is itself evaluated as the Xh component solids carried by a
mixture of the liquid plus Xp and Xf solids, and so on. Furthermore, the volume
concentration of each solids component is evaluated within the mixture containing
only those solids finer than that component. With the exception of the coarsest
fraction Xs, these component volume concentrations are generally larger than the
simple volume concentration of that component within the total slurry, since the
volume of its particular mixture does not include the coarser solids.
The Carrier Fluid Contribution
The fine (Xf) solids interact with the carrying liquid (usually water of the specific
gravity Sl), thereby altering its density and viscosity, but the resulting carrier fluid
mixture is assumed to retain the characteristics of a Newtonian fluid.
The volume concentration of the Xf solids within the mixture of the liquid and Xf
solids (i.e., the carrier fluid) and the specific gravity of the carrier fluid are expressed
as follows:
X f Cv
Cvf = ð5:13Þ
1 - C v ð1 - X f Þ
X f C v ð Ss - S l Þ
Sf = Sl þ ð5:14Þ
1 - C v ð1 - X f Þ
V 2m
i f = Sf f f ð5:16Þ
2gD
X p Cv
Cvp = ð5:17Þ
1 - Cv 1 - X f - X p
X f þ X p C v ð Ss - Sl Þ
Sfp = Sl þ ð5:18Þ
1 - Cv 1 - X f - X p
The excess gradient of the pseudo-homogeneous fraction within the carrier fluid
is calculated according to the equivalent liquid model (ELM) described in Sect. 5.2
applied to the Xp solids conveyed within the carrier fluid of relative density (Sf),
The values of the correlation constants of A″f = 1.0 and A″p = 0.5 are
recommended, or they may be calibrated using test data for the slurry of interest,
when available.
The Heterogeneous Contribution
The volume concentration of the heterogeneous (Xh) fraction within a mixture of the
carrier fluid plus Xp and Xh fractions, and the specific gravity of the combined
mixture are expressed as follows:
X h Cv
Cvh = ð5:22Þ
1 - Cv 1 - X f - X p - X h
134 5 Settling Slurry Flow
X f þ X p þ X h C v ð Ss - Sl Þ
Sfph = Sl þ ð5:23Þ
1 - Cv 1 - X f - X p - X h
where M = 1.0, μs may vary from 0.44 to 0.50 according to experience, and the
reference velocity (V50,h) is calculated according to Eq. (5.8), adjusted for a relative
carrier density of Sfp, and using the mass median particle diameter of the Xh fraction
(i.e., d50h, rather than the d50 of the entire slurry; for definition of d50h, see Fig. 5.13):
d0:35 Ss - Sfp
V 50,h = 44:1 50h
ð5:26Þ
ν0:25
r 1:65
where
νf μ
νr = = f ð5:27Þ
νw μw Sf
The empirical parameter C″ accounts for the reduction in Δih resulting from the
inter-granular support of the Xh particles by the finer fractions. This reduction
depends on the volume of the finer fractions, and C″ varies from a minimum value
at the deposition limit velocity to no effect at or above the velocity at full suspension,
where C″ = 1:
V nc
00 00 00 100,s - V m
C = 1 - Cf X f þ Cp X p ð5:28Þ
V 100,s - V sm,h
where Vsm,h is the deposition limit velocity for a d50h-sized grain, C″ = 1 for any
Vm > V100,s, and V100,s is the velocity at full suspension for a ds-sized grain
(ds = 0.015D). The velocity V100,s can be determined as the flow velocity at which
the pipe system curves for heterogeneous and pseudo-homogeneous flow of the ds
slurry intersect:
5.5 Broadly Graded Settling Slurry Flow 135
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
μs
V 100,s = 3
g D V 50,s ð5:29Þ
A00 f f,100,s
C vs = X s Cv ð5:30Þ
and the specific gravity of the combined mixture is simply the slurry specific gravity
(Sm).
The excess gradient of the fully stratified component is calculated according to
the Vsm model described in Sect. 5.3 applied to the Xs solids conveyed within the
three-component heterogeneous mixture of relative density Sfph,
0:25
00
V sm,s
Δis = B 2μs C vs Ss- Sfph ð5:31Þ
Vm
where the model reference velocity (Vsm,s) is the deposition limit velocity for the ds-
sized grain (ds = 0.015D).
As with C″, the parameter B″ accounts for the reduction in Δis resulting from the
inter-granular support of the Xs particles by the finer fractions, depends on the
volume of the finer fractions, and varies from a minimum value at the deposition
limit velocity to no effect at or above the full-suspension velocity, where B″ = 1:
V nB
00 100,s - V m
B =1- B00f Xf þ B00p Xp þ B00h Xh ð5:32Þ
V 100,s - V sm,s
Note that the deposition limit velocity for the total flow (all fractions) can be
higher than Vsm,s and, in general, Vsm ≈ Vsm,h > Vsm,s. Based on laboratory testing by
Visintainer et al. as described below, default values for the correlation constants of
B″f, = 1.0, B″p, = 1.0, B″h = 0.5, and nB = 0.5 are recommended. Alternatively,
these constants may be calibrated using test data for the slurry of interest, when
available.
136 5 Settling Slurry Flow
The benign effect of a broad particle-size distribution is most readily observed in the
flow of bimodal slurries composed of a fraction of very coarse grains with an
additive of a finer fraction. The experiments by Visintainer et al. (2017) contained
tests with various bimodal slurries, from which combinations of (Xs + Xf), (Xs + Xp),
(Xs + Xh), and (Xh + Xp) were the most interesting in terms of a possible friction
reducing effect that the finer fraction might have on the coarser solids, as expressed
by the empirical parameters C″ (Eq. (5.28)) and B″ (Eq. (5.32)) in the 4CM.
Interestingly, the pseudo-homogeneous component had the most pronounced
effect of all three components when mixed with Xs. In fact, the addition of the Xp
fraction reduced the overall friction so that the bimodal slurry flow losses were
smaller than those in the flow of Xs slurry alone (Fig. 5.16). The Xp component was
the most effective in reducing the mechanical friction between the sliding bed and
the pipe wall. The addition of Xp component to the Xh slurry did not show the same
friction-reducing effect, although the total frictional im did remain virtually
unchanged after adding the Xp fraction. This is because mechanical friction is of
5.5 Broadly Graded Settling Slurry Flow 137
0.30
Cvd = 0.379
0.25 4CM prediction
0.20
im (m/m)
0.15
0.10
0.05
D = 203mm
0.00
0 2 4 6 8
Vm (m/s)
1.0
0.9 40 m m 0.015D
Xp max dh
0.8
0.7
0.6
% passing
0.5
0.4
0.3
0.2
0.1
0.0
0.01 0.1 1 10
d (mm)
Fig. 5.14 Dimensionless pressure gradient (above) measured and predicted by the 4CM for a broad
particle size distribution slurry (below) in a 203-mm pipe. Experimental data from Visintainer et al.
(2017)
138 5 Settling Slurry Flow
0.05
0.00
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
(Sm - 1) (-)
0.40
Xs, Cvd = 0.154
0.35
Xs + Xp, Cvd = 0.154 + 0.153 = 0.307
0.30
0.25
im , if (-)
0.20
0.15
0.10
0.05 D = 203mm
0.00
0 2 4 6 8
Vm (m/s)
Fig. 5.16 Dimensionless pressure gradient for individual and bimodal slurries in a 0.203 m pipe.
Points show experimental data, solid lines the 4CM prediction. (Matoušek et al. 2019)
less consequence with a partially stratified Xh slurry than it is with a fully stratified Xs
slurry. Interested readers may refer to Matoušek et al. (2019) for more details and
comparisons based on these experimental results.
Various combinations of solids fractions in bimodal slurries were also tested in
other laboratories, and the reported results agreed with the trends observed in the
GIW tests. Maciejewski et al. (1993) found that fine tar sand with d50 ≈ 150 μm was
effective in reducing the solids effect of very coarse (~100 mm) rock in slurry flow
through a 489-mm pipe.
5.6 Flow Over a Stationary Bed 139
Another type of stratified flow that may be encountered is flow above a stationary
bed of solids. This type of operation is generally uneconomic and is thus seldom a
deliberate design choice. It is, however, sometimes found in existing pipelines, and
although less popular now, has been considered a viable option in the past, especially
where the particles are very abrasive. However, such operation is not without its
dangers, as outlined below.
Steady flow over a stationary bed has been studied experimentally in pressurized
pipes of laboratory loops with the primary aim of acquiring experimental data for the
development of formulae for interfacial (top-of-bed) resistance and for transport at
high bed-shear.
An appropriate predictive technique for such flows in pipes is a version of a
layered model that considers the stationary deposit instead of the sliding bed as the
lower layer in the layered structure of the flow (Fig. 5.17). This model can be a
component of a unified layered model that identifies whether the bed slides or not
and solves the flow for both cases (Matoušek et al. 2018a). The model predicts the
hydraulic gradient, the thickness of the deposit, and the solids distribution above the
deposit.
For some preliminary estimates of the frictional pressure drop above a stationary
deposit, predictions by the 4CM (which is normally used at Vm > Vsm) are also
useful.
Ow
Aa us D
va
Cb
yb
Ob
Fig. 5.17 Schematic sketch of internal structure of settling slurry flow above a deposit
140 5 Settling Slurry Flow
In long slurry pipelines operating at velocities below the deposition limit, and which
are fed by slurry of variable density, the density fluctuations experienced at the
pipeline inlet may result in the development of density waves along the length of the
pipeline (Matoušek 1996, 1997). Figure 5.18 shows the results of density measure-
ments at three locations along a 650-mm pipeline fed by a dredge with an aqueous
slurry of medium sand (median size 0.25 mm). The measurements were continuous
and took place at a dredge near the pipeline inlet and at two booster stations at
distances of 2.25 km and 6.90 km, respectively, downstream of the dredge. Further-
more, the mixture velocity was continuously measured at one of the boosters. This
signal is plotted in the figure as well.
The shapes of the density signals suggested that short time-scale fluctuations
generated at the inlet decayed (due to turbulent diffusion) while long time-scale
fluctuations tended to amplify, particularly if the mixture velocity dropped far below
the deposition limit (which was observed to be approximately 3.4 m/s). The largest
peaks, developed at the lowest throughput velocities in the pipeline, were able to
grow into very dense masses if the velocity remained low for a long time
(Fig. 5.19b). If the mixture velocity was maintained above the deposition limit,
then no amplification of the density waves occurred (Fig. 5.19a). The presence of a
stationary bed leads to an interaction between the bed and the suspension flow above
the bed and results in mass exchange (due to sedimentation and erosion) between the
stationary deposit and the flow above the deposit. For more on the development of
3.20
2.70
13:00:00 14:00:00 15:00:00 16:00:00 17:00:00 18:00:00
flow meter location: JAGERSPLAS (1886 m)
Slurry density Slurry density [kg/m3]
1250
1000
12:30:00 13:30:00 14:30:00 15:30:00 16:30:00 17:30:00
density meter location: JAGERSPLAS (1886 m)
1500
1250
1000
12:40:00 13:40:00 14:40:00 15:40:00 16:40:00 17:40:00
density meter location: DUINJAGER (6538 m)
Slurry density
1500
1250
1000
13:00:00 14:00:00 15:00:00 16:00:00 17:00:00 18:00:00
Time
Fig. 5.18 Development of density waves measured in a long pipeline transporting medium sand
slurry fed by dredge. (Matoušek 1996)
5.7 Review of CFD Modeling 141
Fig. 5.19 Development of density waves at different throughput velocities for measurements as in
Fig. 5.18. (a) at velocity above the deposition limit, (b) at velocity below the deposition limit.
(Matoušek 2001)
density waves and associated mechanisms, see Matoušek (2001), Miedema et al.
(2003), Talmon (1999), Talmon et al. (2007), and de Hoog et al. (2021, 2022).
In a long transportation system with booster pumps along the line, the density
waves tend to overload the booster pump drives. If very dense, the waves may pose a
danger of pipeline blocking (e.g., in elbows from horizontal to vertical). However,
from the field measurements discussed above, the waves did not appear to signifi-
cantly increase the frictional head losses—or, hence, the total pipeline resistance.
complexity; and at their most complex involve the use of direct numerical simulation
(DNS),1 which solves the entire fluid motion using the Navier-Stokes equations
directly, combined with discrete element modeling (DEM), where the motion of
individual particles is computed, along with their effect on the fluid. Highly coupled
models using complex turbulence models and discrete particles, such as DNS/DEM,
are extremely expensive computationally. Even though the number of particles
modeled may exceed hundreds of millions, this is still too small for practical use
in many hydraulic conveying applications. In addition, the inner time steps required
to model particle-particle collisions with DEM are extremely small (i.e., 10-6 s),
further increasing the computational requirements. Finally, DEM simulations do not
scale well in parallel computing, and it is difficult to accurately model the drag and
lift forces on the particles in dense flows, where the coefficients are altered due to the
close proximity of the particles to each other.
In cases where interactions at or very close to the wall must be considered, such as
with friction losses or erosion modeling, the wall treatment plays a pivotal role in
producing accurate predictions. Some simulations rely on wall functions to avoid the
necessity of resolving the boundary layer near the wall with very thin divisions of the
computational mesh, thereby reducing the computational expense. However, the
presence of particles can influence the physics at play within the boundary layer; in
order to produce accurate predictions, the model may need to account for these
effects using solids modified wall functions.
In order to get the most accurate predictions for friction losses, CFD models must
account for the velocity distribution in the area of shear close to fixed boundaries or
walls, in particular, within the buffer layer between laminar and turbulent flow
beginning near the dimensionless wall distance y+ = 5, (see Sec. 2.5 and
Fig. 2.14). This means the first layer in the computational mesh may be just a few
microns in thickness. For many of both the Eulerian and Lagrangian models, there is
a requirement that the particles be sufficiently smaller than the grid size; hence, the
particles are often too large to be accurately modeled close to the wall. To deal with
larger particles and the associated mesh size limitations, another possibility is to
make the grid size much smaller (for example 20 times smaller) than the particle,
thereby allowing integration of the pressure and shear stress fields over the face of
the particle and eliminating the need for drag coefficients. Unfortunately, this
treatment is extremely computationally expensive and is not viable for the modeling
of industrial sized systems using presently available computing resources.
1
Computational modeling of turbulent flow can take several forms, the difference being in the way
the complex spectrum of macro- and micro-interactions of a turbulent flow are modeled. With
Reynolds Averaged Numerical Simulation (RANS), this complexity is modeled using relatively
simple algebraic equations. At the other extreme, Direct Numerical Simulation (DNS) attempts to
apply the full Navier-Stokes equations to the entire fluid flow (which, understandably, is a very
expensive process). A compromise, Large Eddy Simulation (LES), models only the very smallest of
these interactions using suitable algebraic models, while still attempting to model the larger
turbulent structures.
144 5 Settling Slurry Flow
Case Study 5.2 demonstrates the sensitivity of the flow behavior (and hence the
pumping requirements), of a suspension, to the solids’ particle-size distribution.
Hydraulic transport, especially when under turbulent flow as described here, can
modify the particle-size distribution, through particle-wall interaction and, at suffi-
ciently high solids concentration,2 by particle-particle interaction. The collisions of
these interactions break pieces from the particles, generally rounding them while
generating additional fine material. The severity of this action is a function of both
the solids properties and the turbulence intensity. If the material is very friable or
fragile (e.g., coal, mud stones, laterites, bauxite, nascent crystals), the production of
fine particles through attrition can be extreme. For harder materials (e.g., silica sands
and gravels), it is usually only moderate. Particle attrition can modify the transport of
the solids and the whole process in several ways:
• Inadequate separation. The production of large amounts of fines can devalue a
product through the failure to meet a specified distribution. The increase in fines
also decreases the effectiveness of separation techniques, often requiring second-
ary separation circuits or more advanced, and hence expensive, methods. Failure
2
At very high concentrations, particle attrition is minimal due to the reduced particle movement
(Bhattacharya & Imrie 1986)
5.8 Particle Attrition 145
to separate the fluid adequately can result in storage problems in stockpiles and
with cargo stability during shipping. These latter problems are a result of
increased mobility through liquefaction, which can cause the stockpile or cargo
to shift, resulting in damage or even the capsizing of the cargo ship.
• Reduction in transport pressure gradient. The rounding of the particles and
increase in the fines fraction reduces the transport pressure gradient in two
ways. First, the rounding results in a lower coefficient of sliding friction (μs),
which reduces the contribution from the Xh and Xs fractions to the pressure
gradient. Second, the fines increase the Xf and Xp fractions, thus increasing their
mixture densities and reducing the contributions of the submerged fractions Xh
and Xs to the overall pressure drop.
• Change in flow properties. If the solids’ surface chemistry is suitable, the
produced fines will be rheologically active. This can affect dewatering systems
through the action of occluded water in the self-agglomerating flocs (i.e., clusters
of fine particles). In extreme cases, it can change the carrier fluid from water to a
non-Newtonian mixture. In tests conducted to establish the effect of particle
attrition on coarse coal slurries (Elliot and Gliddon 1970), a conventional turbu-
lent settling suspension became a non-settling hybrid suspension being conveyed
in laminar flow. The implications of these results are discussed in Chap. 6.
• Pipe wear (entry region). Particle rounding usually produces a less abrasive
slurry, which reduces pipe wear. Unfortunately, this amelioration is only signif-
icant after the first couple of kilometers of transport. Before that there is still
sufficient oxygen in the slurry to produce a corrosive/erosive environment which,
together with the more abrasive (pre-attrition) particles, may result in significantly
higher pipe wear in the entry region.
• Effects of pipe geometry, pump passages, and casings. Particle attrition is dra-
matically increased by changes in pipe geometry through bends and fittings, and
as particles pass through pump passages and casings. Such changes in flow
should be minimized, especially for fragile materials, by keeping the piping as
uniform and straight as possible and bypassing the pump via lock hoppers or
similar devices. Generally, these measures are not possible in loop testing, where
the particles continually traverse the loop passing through the pump during each
circuit. Under such conditions, it is essential that particle size distributions (PSDs)
are sampled frequently during testing and solids replaced whenever their attrition
exceeds that which can be accounted for in the analysis.
Where particle attrition is a known issue, it is possible to measure the severity of
the particle breakage and then to model it numerically using models similar to those
used for comminution in ball mills (e.g., Austin and Luckie 1971) and a good review
of this and other subjects for attrition in pipelines is given by Gillies (1991).
Unfortunately, such approaches require the establishment of a matrix of breakage
functions, which are sensitive to both material and flow. As a result, the modeling
produces rather specific results.
146 5 Settling Slurry Flow
A simpler, and often more useful, approach is to measure the attrition directly. As
mentioned, the multiple transits through pump passages, along with the number of
bends in pipe loops, makes pipe test loops a poor candidate for attrition measure-
ment; multiple tests with different length pipes are required, so that the effect of the
pumps and fittings can be separated from that in the pipe. Even when such pre-
cautions are taken, the results can still contain a degree of uncertainty, as fractures
created during multiple passages through the pump may not occur in once-through
systems.
A toroidal wheel stand (Jacobs 1991; Traynis 1977), comprising a pipe bent to
form the periphery of a large vertical wheel that is rotated about a horizontal axis, can
overcome many of these problems. The pipe is two-thirds filled with a slug of the
slurry of interest and rotated at a particular speed, such that the peripheral speed of
the wheel matches the desired transport velocity. The slug of slurry tends to remain
stationary during this process, although it is no longer symmetrical about the vertical
axis, as it is dragged somewhat in the direction of rotation. Although this is only an
analog to actual pipeline flow, and the flow is more complicated, as described by
Adedeji et al. (2022), a reasonable estimate of the conveying gradient can be
obtained by monitoring the height of the two ends of the slug. The flow of the slurry
over the moving pipe wall is a reasonable analog of that undergoing transport in a
pipe. By running the wheel stand for a period equal to the transport time of the target
system, the extent of the particle attrition can then be measured by comparing the
slurry’s PSD before and after the test.
The transport system from Case Study 4.2 is supposed to convey sand over a
distance of 350 m. To finalize the pipe size selection started in Case Study 4.2 and
evaluate pipeline energy requirements, slurry flow parameters in pipes of the three
candidate sizes must be determined and analyzed. Laboratory tests were carried
out with the medium sand under consideration (Ss = 2.65; d15 = 0.33 mm;
d50 = 0.47 mm; d85 = 0.57 mm) in a 103-mm (4-inch) pipe, and the results calibrated
the V50 friction loss model for heterogeneous flow (Sect. 5.4) using the technique
shown in Fig. 5.10. The calibration gave M = 0.97 and V50 = 2.98 m/s (for
μs = 0.48).
(a) Preparation of Friction Loss Model
For the evaluated medium sand, the d50/D < 0.015 in all three candidate
pipes, and therefore a heterogeneous flow model is appropriate to predict pipe
flow characteristics. The V50 model assumes M = 1.0 for natural distributions of
sands and gravels, and it is in excellent agreement with the calibrated value from
5.9 Case Studies 147
the lab experiment. The model reference velocity (V50) is calculated by Eq. (5.8)
using the properties of the solids and liquid under consideration:
d0:35 Ss - Sf 0:000470:35 1:652
V 50 = 44:1 50
= 44:1 = 3:02 m=s:
ν0:25
r 1:65 10:25 1:65
This value is also very similar to that from the experimental calibration. Note
that Eq. (5.8) considers V50 to be independent of D and Cvd. Hence, V50 becomes
a constant in our analysis of the sand slurry flow of various Cvd in pipes of
various sizes. The analysis will employ the predictive V50 model to evaluate flow
conditions in the three candidate slurry pipes.
(b) Determination of Minimum Velocity (Vmin)
For a safe and stable operation of a pipe of a certain diameter (D) transporting
slurry of a certain delivered concentration (Cvd), the operating velocity (Vm)
must exceed the deposition limit velocity (Vsm) (as evaluated in Case Study 4.2).
It must also exceed the minimum velocity (Vmin), which is the velocity at the
minimum on a pipe characteristic curve. The pipe curve is predicted by a friction
loss model, so the same model must also predict the minimum velocity for the
curve. The V50 model can be used to predict Vmin, as shown in Eq. (5.10),
assuming that the effect of Vmin on ff can be neglected. Inserting the values of
solids and liquid properties, Eq. (5.10) gives
1 1=3
C ðS - S ÞgDμ V M M Mþ2
V min = vd s 2Sf f f s 50 = Cvd 1:6529:81D0:483:02
20:998f
f f
1=3
C vd D
= 11:77 f
f
For the range of pipe sizes (D) and solids concentrations (Cvd) selected in
Case Study 4.2, and for the estimated roughness of a new commercial steel pipe
(ε = 45 μm), values of Vmin from this equation are summarized in the table
below. Furthermore, the table relates Vmin and Vsm with the operating velocity
Vm,500 (to ensure the solids discharge Ms = 500 tonnes/hour as required in Case
Study 4.2). It does this by showing values of the velocity ratios Vm,500/Vmin and
Vm,500/Vsm. By comparing these values of the velocity ratios, we can select the
best available combinations of D and Cvd based on the following criteria, which
must be satisfied at the same time. First, Vm,500/Vmin > 1.1 (estimated value) to
avoid operation at or below the minimum friction loss flow rate where the
intersection between the pump and system characteristic curves becomes unsta-
ble. Second, Vm,500/Vsm > 1 to avoid a stationary deposit in the pipe, while at the
same time, being held as low as possible to avoid excessive friction loss
and wear.
148 5 Settling Slurry Flow
They are plotted together with their characteristic velocities (Vsm, Vmin, and
Vm,500) in Fig. 5.20.
The dimensionless pressure gradient im at the operating velocity Vm,500 is
given in Table 5.2. This table also quantifies other friction loss–related param-
eters of the sand slurry transport at Vm,500, namely the specific energy consump-
tion SEC (kWh/tonnes-km) by Eq. 3.10, SEC = 2:73 SsiCmvd = 1:03 Cimvd , and the
head Hm,frict (in m slurry) required to overcome the friction losses in the 350 m
long pipeline:
It follows from the structure of the V50 model that the proportion of solids
M
transported as contact load can be estimated as CCvdc = 0:5 VV50m . If Vm = V50, then
Cc/Cvd = 0.5. In other words, half of the volume of solids is transported as the
contact load (by definition). For the conditions discussed here, CCvdc = 0:5 V3:02
m,500
0.25
Vsm 8" pipe, Cvd
Cvd = 0.35
Vmin Cvd = 0.25
10" pipe, Cvd
Vm_500 Cvd = 0.15
12" pipe, Cvd
0.20
im (-)
0.15
0.10
0.05
2 3 4 5 6
Vm (m/s)
Fig. 5.20 Pipe friction loss curves for the 8-inch pipe at Cvd = 0.35, the 10-inch pipe at Cvd = 0.25,
and the 12-inch pipe at Cvd = 0.15
Table 5.2 Flow conditions for sand slurry transport in the selected pipes at Ms = 500 tonnes/hour
SEC Hm,frict
D (mm)/Cvd (–) Vm,500 (m/s) im (–) (kWh/tonne-km) (m slurry)
202.7/0.35 4.64 0.172 0.505 38.1
254.5/0.25 4.12 0.122 0.502 30.2
304.8/0.15 4.79 0.091 0.622 25.4
contributes to the contact load, and the rest is suspended in the carrying liquid at
the operating velocity in any of the three pipes. This information may be useful,
e.g., for an evaluation of wear in the pipeline.
(d) Evaluation of the Particle Size Effect—Heterogeneous Flow of Coarse Slurry
It is of interest to evaluate how the designed system would perform if a
considerably coarser sand/gravel had to be conveyed at the same solids dis-
charge. Let us consider a coarse fraction (Ss = 2.65, d15 = 1.0 mm,
d50 = 3.0 mm, d85 = 5.0 mm) that is still transported in heterogeneous slurry
flow (the d50/D < 0.015 in all three pipes, although with a very little margin in
the case of the smallest pipe size). The same procedure is followed as for the
transport of the medium sand (steps (a) to (c)) with the following results.
The reference velocity V50 = 5.78 m/s, which is obviously higher than for the
medium sand. The deposition limit velocity (Vsm) is calculated using Eq. (4.14)
with μs = 0.48, instead of Eq. (4.15), which was good for the medium sand.
Although the particles are larger, Vsm is lower than for the sand, but that does not
mean that the operating velocity Vm,500 can be lowered as well. Conversely,
150 5 Settling Slurry Flow
Table 5.3 Flow conditions for transport of coarse solids at Ms = 500 tonnes/hour
D (mm)/Cvd Vm,500/ Vm,500/ Vm,500 im SEC (kWh/tonne- Hm,frict
(-) Vmin Vsm (m/s) (-) km) (m slurry)
202.7 / 0.30 1.20 2.24 5.41 0.237 0.813 55.5
254.5 / 0.20 1.20 1.82 5.15 0.165 0.849 43.4
304.8 / 0.15 1.14 1.49 4.79 0.125 0.857 35.0
shapes of the predicted flow curves indicate that the required criteria for the
velocity ratios (see step (b)) are met only if Cvd is reduced by 0.05 for the 8-inch
pipe (Cvd from 0.35 to 0.30) and for the 10-inch pipe (Cvd from 0.25 to 0.20). A
stable operation would not be assured if Cvd was maintained because
Vm,500 < Vmin. As a result of the Cvd reduction, the Vm,500 becomes about 1 m/
s higher than for the medium sand. This produces a more than generous margin
against deposition. It also contributes to considerably higher friction losses
(almost 50% higher than for the medium sand) and an even greater increase in
SEC. In the largest pipe, the margin remains reasonable, but SEC is high anyway
due to the originally low Cvd of 0.15. All results are summarized in Table 5.3.
It is expected that the coarse slurry is more stratified and that more solids
contribute to contact
load
than for the previously evaluated medium sand.
Indeed, CCvdc = 0:5 V5:78
m,500
results in slightly more than half of the transported
coarse solids contributing to the contact load in all three pipes. This, combined
with the higher Vm,500, contributes to presumably higher wear on the pipeline
than if the medium sand would be conveyed in the system.
The system design procedure progresses with a selection of a suitable pump
for each of the three possible pipes and with an evaluation of the energy
requirements and efficiency of the entire system. This procedure is continued
in Chap. 9.
This case study is concerned with the effect of broadly graded solids on the pipeline
transport of settling slurries. Perhaps the best way of demonstrating this effect, and
the behavior of broadly graded slurry flow in general, is to consider the example of
conditions tested experimentally in a laboratory pipe loop (based on Visintainer et al.
2022). The tested solids have a density (ρs) of 2650 kg/m3 and contain particles
spanning a broad range of sizes from tens of microns to almost 20 mm. Slurry flow of
these solids will be evaluated and compared with the flow of a considerably more
narrowly graded solids of a very similar median size (d50), also tested in the lab. The
evaluation will be carried out for transport in a 20-inch pipe (D = 489.0 mm). The
effect of the pipe size on the broadly graded slurry flow will be considered by also
including a prediction of flow in an 8-inch pipe (D = 202.7 mm).
5.9 Case Studies 151
100
90
80
Cummulative passing (%)
70
60
50
40 A
30
B
20
10
0
0.01 0.1 1 10 100
d (mm)
Fig. 5.21 Particle size distribution curves for the evaluated broadly graded Slurry A and the more
narrowly graded Slurry B. Dashed lines denote fraction thresholds (red): Xf to Xp, (blue): Xp to Xh,
(green): Xh to Xs
(a) Settling Slurry Flow of the Broadly Graded Slurry in a 20-Inch Pipe
The particle size distribution curve of the broadly graded slurry (Slurry A in
Fig. 5.21) gives d50 = 3.3 mm, d85 = 13.9 mm and, for the purposes of the
4-Component friction loss model (4CM), Xf = 0.07, Xp = 0.24, Xh = 0.37, and
Xs = 0.32. Therefore, all four components considered by the 4CM are present in
the solids distribution, although the presence of the finest component (Xf) is
minor. The heterogeneous component (Xh) is the largest (and its d50h = 3.1 mm),
but the contributions of the coarsest, stratified, component Xs (dmax = 19 mm)
and of the pseudo-homogeneous component (Xp) are also important.
We consider flow of this Slurry A at a temperature of 23 °C in a 20-inch pipe
at the delivered concentration of Cvd = 0.21, noting that the laboratory pipe is
made of steel and has a very low roughness of ε = 1.5 μm. Ignoring the
components for now, let us look only at the median size (d50) for the slurry
flow characterization. The d50/D = 3.3/489 = 0.007, which is less than the upper
threshold value of 0.015 for the heterogeneous slurry. Thus, the V50 model
(M = 1.0, μs = 0.48) might be considered for the flow prediction. This model
gives a Vmin = 6.30 m/s (Eq. (5.10)), which is substantially above the deposition
limit velocity of Vsm = 4.29 m/s. (This value is obtained with Eq. (4.15),
which gives a slightly lower value than Eq. (4.14) for μs = 0.48.) The pipe
flow curve (the red line in Fig. 5.22) shows that the gradient im = 0.121 at
Vm = 6.9 m/s, which is selected as a potentially suitable operating velocity
(based on Vm = 1.1 Vmin).
152 5 Settling Slurry Flow
0.20
CCvd
vd
= 0.21
V50 model
4CM
0.15 water curve
im (-)
0.10
0.05
0.00
2 3 4 5 6 7 8 9 10
Vm (m/s)
Fig. 5.22 Pipe flow curves for a broadly graded slurry in a 20-inch pipe
the 4CM starts to predict a Vmin slightly above Vsm. For instance, if Cvd = 0.35,
then Vmin is 7.5 m/s by the V50 model and 4.5 m/s by the 4CM. This suggests the
operating velocities of 8.3 m/s in the first case, and 5.0 m/s in the other, a
difference that is 40% larger than for Cvd = 0.21. However, the difference in the
corresponding pressure gradients remains almost the same.
Based on the above analysis, we can initially conclude that the friction loss
models predict lower losses for broadly graded solids than for narrowly graded
solids of the same median size (d50) at velocities not too far above Vsm and Vmin
(i.e., at velocities potentially attractive as the operating velocity). It is of interest
to verify whether this trend is real, so in the next section, we examine experi-
mental data for a more narrowly graded slurry flow.
(b) Comparison with More Narrowly Graded (Dominantly Heterogeneous) Solids
with a Very Similar d50 in a 20-Inch Pipe
A slurry with more narrowly graded solids (Slurry B in Fig. 5.21) suitable for
a comparison with the broadly graded Slurry A was tested in the same 20-inch
pipe loop. Its properties were as follows: density (ρs) was 2650 kg/m3,
d50 = 3.5 mm, d85 = 7.2 mm, d50h = 3.55 mm, dmax = 13 mm, Xf = 0.05,
Xp = 0.10, Xh = 0.71, Xs = 0.14. This narrowly graded slurry is not solely
one-component, but the Xh component dominates with 71% of all particles.
Flow of Slurry B was tested at a temperature of 25 °C and at the delivered
concentration Cvd = 0.11 in a 20-inch pipe. The V50 model is more appropriate
for this slurry than for the previously evaluated and broader Slurry A. It predicts
Vmin = 5.10 m/s, which is higher than the deposition limit velocity
Vsm = 4.29 m/s. (Both Eq. (4.15) and Eq. (4.14), for μs = 0.48, give the same
value, and it agrees with the visually observed 4.3 m/s.) The V50 model ade-
quately describes the observed flow curve. The 4CM does not predict Vmin > Vsm,
0.20
Cvd = 0.11
Cvd
V50 model
4CM
0.15 water curve
im (-)
0.10
0.05
0.00
2 3 4 5 6 7 8 9 10
Vm (m/s)
Fig. 5.23 Pipe flow curves for a less broadly graded slurry in a 20-inch pipe
154 5 Settling Slurry Flow
and it agrees with the observed shape of the pipe flow curve as well as, if not
slightly better than, that of the V50 model (Fig. 5.23). This agreement between
the models is much closer than before, as might be expected. Over the range of
appropriate operating velocities, the difference in the predictions of im is less
than 10%.
This result demonstrated that the 4CM can be successfully used for slurries
where one component dominates, because it effectively reduces to one of its
sub-models when handling a one-component slurry (e.g., it reduces to the V50
model if Xh = 1, or to the Vsm model if Xs = 1).
(c) The Effect of Pipe Size on the Flow of the Broadly Graded Slurry (8-Inch Pipe)
If a broadly graded slurry is considered for transport in a pipe of a smaller
diameter, then the slurry becomes effectively “coarser” due to the increase in the
d/D threshold between Xh and Xs, moving a portion of particles from the Xh
component into the Xs component. For an evaluation of the transport of the
broadly graded Slurry A in an 8-inch pipe (steel, ε = 1.5 μm) instead of a 20-inch
pipe, this criterion disqualifies the V50 model from use (d50/D = 3.3/
202.7 = 0.0163) and modifies Xh and Xs for the 4CM (Xh = 0.19 and
Xs = 0.50 instead of Xh = 0.37 and Xs = 0.32). Furthermore, d50h shifts from
3.1 mm to 1.4 mm. The “coarsening” effect of the slurry on the pressure gradient
at the operating velocity is also not negligible. If Vm = 1.1Vsm = 2.6 m/s
(Vsm = 2.36 m/s by Eq. (4.14) for μs = 0.48 giving a lower value than
Eq. (4.15) in the 8-inch pipe), then im = 0.130 for the modified Xh, Xs, d50h,
compared with 0.121 if the modification is ignored.
dredge. The dredge is connected to a discharge line that carries the coarse slurry to a
shore disposal area. The discharge pipe has an inner diameter D = 489 mm and a
roughness typical for commercial steel pipe, ε = 45 μm. The cutter head and the pipe
suction mouth feed the system with slurry of the desired delivered concentration
Cvd = 0.215. The slurry flow rate is measured onboard the dredge, and its typical
value is 1.13 m3/s, corresponding to the slurry flow velocity Vm = 6.0 m/s. The
solids density (ρs) = 2650 kg/m3, and the carrying liquid is seawater of a temperature
of 15 °C, a density of 1026 kg/m3, and a dynamic viscosity of 0.00122 Pa.s.
Let us first assume that a clean, narrowly graded gravel (characteristic size
11 mm) is dredged. (Because it is clean, the gravel does not contain particles fine
enough to be suspended in the turbulent flow.) The slurry flow is considered fully
stratified; thus, the Vsm model is used to estimate the friction loss. The deposition
limit velocity Vsm ≈ 3.2 m/s (Eq. (4.14) for μs = 0.48). For the flow conditions under
consideration, the dimensionless pressure gradient (meters of standard water/unit
length of pipe) for the seawater
V 2m 1:026ð0:0126Þð6:0Þ2
i f = Sf f f = = 0:0485
2gD 2ð9:81Þð0:489Þ
This is quite a high value—approximately three times larger than im for the
broadly graded slurry flow of the same Vm, and very similar Cvd in a pipe of the
same D that was evaluated in Case Study 5.2(a).
However, the natural gravel dredged from the seabed is obviously not mono-
sized, and a sample taken from the discharge line specifies its particle size distribu-
tion as follows: d50 = 10.7 mm; d85 = 16.5 mm; d50h = 5.3 mm; dmax = 19.5 mm;
Xf = 0.12; Xp = 0.03; Xh = 0.16; Xs = 0.69. It follows that the original estimate of
the friction loss must be modified by using the 4CM friction loss model for broadly
graded solids.
Figure 5.24 shows the pipe flow curve as calculated by the 4CM. It exhibits Vmin
close to 4.5 m/s and confirms that the current operating velocity of 6.0 m/s is
reasonable, although slightly lower velocities might cause the transport to be slightly
less energy-consuming and significantly less erosive to the pipeline.3 Following the
4CM prediction, Vm = 1.1Vmin = 4.8 m/s and the corresponding gradient im = 0.206.
At the current operating velocity Vm = 6.0 m/s, the im = 0.215, not significantly
larger than at 4.8 m/s. It is, however, 34% less than predicted by the Vsm model,
3
Pipe wear is typically proportional to the cube of the conveying velocity or thereabouts.
156 5 Settling Slurry Flow
0.40
4CM, Cvd = 0.215
0.35
water curve
0.30
0.25
im (-)
0.20
0.15
0.10
0.05
0.00
2 3 4 5 6 7 8
Vm (m/s)
Fig. 5.24 Pipe flow curve for natural gravel slurry in a 20-inch pipe
which assumed all particles in contact load (the equation above gave im = 0.335 at
Vm = 6.0 m/s for Cvd = 0.215).
It is worth noting that flow curves of coarse slurries in large pipes tend to be
shallow. Therefore, a small change in the required head may produce a relatively
large change in slurry velocity and discharge. This fact may be of importance in an
evaluation of operational stability when considering pipe-pump transport systems, as
discussed in Chap. 10.
References
O.E. Adedeji, L. Zhang, N.R. Sarker, D.E.S. Breakey, R.S. Sanders, Characterization of the
hydrodynamics within a toroid wear tester. Can. J. Chem. Eng. 100, 1941–1953 (2022)
M.J. Andrews, P.J. O'Rourke, The multiphase particle-in-cell (mp-pic) method for dense particulate
flows. Int. J. Multiphase Flow. 22, 379–402 (1996)
L.G. Austin, P.T. Luckie, Methods for determination of breakage distribution parameters. Powder.
Tech. 5, 215–222 (1971)
A. Bhattacharya, I. Imrie, Hydraulic transport of solids in pipes. Presented at the 10th international
conference on the hydraulic transport of solids in pipes, Innsbruck, 1986
E.D. Cristea, P. Conti, Hybrid Eulerian multiphase–dense discrete phase model approach for
numerical simulation of dense particle-laden turbulent flows within vertical multi-stage cyclone
heat exchanger. Presented at the ASME 5th joint US-European fluids engineering summer
conference, FEDSM, Montreal, 2018
References 157
E. De Hoog, A. Talmon, C. van Rhee, Unstable transients affecting flow assurance during hydraulic
transportation of granular two-phase slurries. J. Hydraul. Eng. 147(9), 04021029 (2021)
E. De Hoog, A. Talmon, C. van Rhee, Pipeline design – Density wave amplification and slurry
dynamics. Terra et Aqua, IADC 166, 15–25 (2022)
R. Durand, Transport hydraulique des matériaux solides en conduite, études expérimentales pour
les cendres de la central arrighi. La Houille Blanche 6, 3384–3393 (1951)
R. Durand, Basic relationships of the transportation of solids in pipes - Experimental research.
Presented at the international association for hydro-environment engineering and research
congress, Minneapolis, 1953
R. Durand, E. Condolios, Donnees techniques sur le refoulement hydraulique des matériaux solides
en conduite. Journess de l’Hydraulique 2, 27–55 (1952)
K. Ekambara, R.S. Sanders, K. Nandakumar, J.H. Masilyah, Hydrodynamic simulation of hori-
zontal slurry pipeline flow using ANSYS-CFX. Ind. Eng. Chem. Res. 48, 8159–8171 (2009)
D.E. Elliot, B.J. Gliddon, Hydraulic transport of coal at high concentrations. Presented at the 1st
international conference on the hydraulic transport of solids in pipes, Cranfield, 1970
R. Gibert, Transport hydraulique et réfoulement des mixtures en conduites. in Annales des Ponts et
Chausées, Mai–Juin 1960, Juil-Août (1960)
R.G. Gillies, Particle size degradation in slurries in slurry handling design of solid liquid systems, in
Slurry Handling, ed. by N. Brown, N. Heywood, (Elsevier Applied Science, London, 1991),
pp. 227–237
R.G. Gillies, C.A. Shook, Modeling high concentration settling slurry flows. Can. J. Chem. Eng. 33,
709–716 (2000)
R. Gillies, K.B. Hill, M.J. McKibbem, C.A. Shook, Solids transport by laminar Newtonian flows.
Powder. Tech. 104, 269–277 (1999)
B.E.A. Jacobs, Additional aspects of slurry systems in design of slurry transport systems, in Design
of Slurry Transport Systems, ed. by B.E.A. Jacobs, (Elsevier Science Publishers Ltd, Barking,
1991)
W. Maciejewski, J. Oxenford, C.A. Shook, Transport of coarse rock with sand and clay slurries.
Presented at the 12th international conference on slurry handling and pipeline transport, Brugge,
1993=
V. Matoušek, Solids transportation in a long pipeline connected with a dredge. Terra et Aqua, IADC
3, 11 (1996)
V. Matoušek, in Flow Mechanism of Sand-Water Mixtures in Pipelines. Dissertation, Delft Uni-
versity of Technology, Delft, 1997
V. Matoušek, On the amplification of density waves in long pipelines connected with a dredge.
Presented at the 16th world dredging congress – WODCON XVI (Kuala Lumpur, 2001)
V. Matoušek, P. Vlasák, Z. Chára, J. Konfršt, Experimental study of hydraulic transport of coarse
basalt. in Proceedings of the Institution of Civil Engineers-Maritime Engineering, vol.
168 (2015), pp. 93–100
V. Matoušek, J. Krupička, M. Kesely, A layered model for inclined pipe flow of settling slurry.
Powder Tech. 333, 317–326 (2018a)
V. Matoušek, R. Visintainer, J. Furlan, A. Sellgren, Threshold criteria for components of predictive
model for pipe flow of broadly-graded slurry. Presented at the ASME 5th joint US-European
fluids engineering summer conference FEDSM, Montreal, 2018b
V. Matoušek, R. Visintainer, J. Furlan, A. Sellgren, Frictional head loss of various bimodal settling
slurry flows in pipe. Presented at the ASME-JSME-KSME joint fluids engineering conference
AJKFLUIDS, San Francisco, 2019
V. Matoušek, R. Visintainer, J. Furlan, A. Sellgren, Pipe-size scale-up of frictional head loss in
settling-slurry flows using predictive models: experimental validation. Presented at the ASME
fluids engineering division summer meeting FEDSM, Orlando, 2020
G.V. Messa, V. Matoušek, Analysis and discussion of two fluid modeling of fully suspended slurry
flow. Powder. Tech. 360, 747–768 (2020)
158 5 Settling Slurry Flow
6.1 Introduction
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 159
R. Visintainer et al., Slurry Transport Using Centrifugal Pumps,
https://doi.org/10.1007/978-3-031-25440-6_6
160 6 Non-Newtonian Slurries and Suspensions
Table 6.1 Interrelated advantages and disadvantages of non-Newtonian slurries and suspensions
Advantages Disadvantages
Very low or no limiting flow rate. High velocities may require very high
pressures.
Can be stopped and restarted with ease. May require high restart pressure if the slurry
is thixotropic, or if there is ongoing chemical
reaction.
Can be pumped in laminar flow. Any coarse particles will stratify in
laminar flow.
Does not readily settle. Difficult to separate solids from fluid using
conventional equipment (e.g., jigs, screens, or
cyclones).
Very high concentrations can be achieved, pro- The slurry viscosity is a very strong function of
viding very low specific energy consumptions. solids concentration.
Low flow rates mean low pipeline wear and Scaling or coating of pipes may be an issue.
particle attrition rates.
Readily pumped by both centrifugal and posi- Efficiency derating may be considerable in
tive displacement pumps. centrifugal pumps.
Heat transfer is low (when heat retention is Heat transfer is low.
desirable).
Fig. 6.1 Typical variation of slurry rheology with various parameters: (a) particle concentration;
(b) particle size; (c) pH
Apart from their rheological properties and constitutive form, non-Newtonian slur-
ries and suspensions can be broadly classified into two types:
162 6 Non-Newtonian Slurries and Suspensions
The majority of mining and mineral processing plants produce and use relatively
large particles (d > 50 μm) that are only slightly affected by inter-particle forces, and
these are typically processed at low concentrations (Cv < 40%) where inter-particle
forces are similarly low. These mixtures can usually be treated as having Newtonian
properties (i.e., the equivalent liquid model) or as a settling slurry, as covered in
Chap. 5. On the other hand, many operations and processes (e.g., dredging, phos-
phate mining, alumina production, and the harvesting of fly ash from electrostatic
precipitators) are characterized by large amounts of very fine particles, typically less
than 20 μm, the surfaces of which are electro-chemically active. At moderate
concentrations, these particles combine with the conveying water to form
non-Newtonian slurries or even pastes that do not readily settle but instead behave
as a single-phase fluid. Because the slurry is no longer Newtonian, its behavior will
differ in detail from that of the Newtonian fluids described in Chap. 2, although the
underlying principles given in Sect. 2.4 will still apply.
Industrial slurries handled by centrifugal pumps are generally visco-plastic (i.e.,
shear thinning) and exhibit yield stress. They are often thixotropic (i.e., sensitive to
shear history), whereby the viscosity of the slurry reduces by some amount after
being sheared, the amount of the reduction being dependent upon both the amplitude
and time of the shearing. Thixotropic slurries will however rebuild either completely
or to some extent once the shearing has finished, the time taken and extent of the
rebuild being a property of the material. This reversible reduction should not be
confused with the dramatic reduction in viscosity that flocculated slurries suffer
when subjected to high shear, such as during the passage through a centrifugal pump.
This latter reduction is due to the destruction of the flocculated network of polymer
chains and is irreversible and is called rheomalaxis. Slurries often exhibit additional
elastic properties (viscoelasticity) if the flow conditions change very rapidly, espe-
cially if they have been flocculated using a polymer.
Both thixotropic and elastic behavior are generally beyond the scope of this text,
but fortunately, most industrial systems are designed on the basis of steady-state
conditions, where time scales are long, and the slurries may be considered inelastic.
Furthermore, with the exception of special cases (such as flow from the underflow of
a thickener to a pump), most industrial flows have passed through at least one pump
6.2 Homogeneous Non-Newtonian Slurries 163
Fig. 6.2 Transport characteristics of thickened tailings flowing through a DN315 pipe. Annotation
shown for the Cw = 0.72 suspension
and have generally suffered sufficient shear to reduce the slurry’s viscosity to an
equilibrium condition, which may then be used for all downstream calculations.1
The viscosity of non-Newtonian slurries (η) is often several orders of magnitude
higher than that of water. Consequently, the Reynolds number, Re = ρmVD/η, is
much lower, and both laminar and turbulent flow can be expected in both laboratory
and industrial settings. Figure 6.2 shows the transport characteristics of a non-
Newtonian slurry that is typical of the mining and mineral industries. The pressure
gradient is a non-linear function of solids concentration in the laminar flow regime,
whereas in turbulent flow, it is relatively independent of this value. Whether the
turbulent curves cross each other, as shown here, or form a steadily increasing
value with solids concentration, depends on the slurries’ rheological constitutive
relationships. Many slurries are too viscous to become turbulent at industrially
1
If the materials are thixotropic, a reduction in viscosity will occur in the suction piping to the
pump; however, since this distance is usually quite short, suction-pipe design based on the viscosity
of the material at exit from the underflow, etc., can be used without substantial error and provides a
conservative design.
164 6 Non-Newtonian Slurries and Suspensions
relevant velocities and pipe sizes, and even if turbulence is reached, they do not
attain the high Reynolds number flows typical of water-based systems. Ulti-
mately, both laminar and turbulent flow can provide viable solutions for
transporting such slurries, and the ability to predict the behavior in both regimes
is required.
For laminar flow, analytical solutions that relate the wall shear stress, or pressure
gradient, with mean velocity or flow rate, can be derived from the rheological model
of choice using the same integral technique outlined in Sect. 2.5. However, before
we can apply this method to the rheological models in Table 2.3,2 it is necessary to
consider the effect of yield stress, if one exists. To begin, consider the stress
distribution inside the pipe shown in Fig. 6.3.
The stress distribution varies linearly with pipe radius, as discussed in Chap. 2,
being zero at the center and reaching a maximum value at the wall, denoted as τ0.
Flow only occurs in the pipe if the wall shear stress (τ0) exceeds the slurry yield
stress (τy). Even after flow develops, regions inside the pipe, where the applied shear
stress is less than the yield stress, remain unsheared. Given the conical stress
distribution in a pipe, this results in a sheared annular flow that carries within it an
unsheared, concentric plug flow. The diameter of this plug is simply:
t un-sheared plug
r
D Dp
2
The Cross model is not considered here due to its similarity in behavior over the range of shear
rates of concern with the simpler power-law fluid, as mentioned in Chap. 2.
6.2
Table 6.2 Relationship between wall shear stress and velocity as a function of radial location and bulk velocity in laminar flow for various non-Newtonian fluid
types
Model u(r, τ0) V(τ0)
Newtonian τ0 R r 2 τ0 D
1- 8μ
2μ R
Ostwald– 1=n
n
1 nþ1
τ0 n D n τ0
de Waele nþ1 k
R 1- Rr n 2 3n þ 1 k
(power
law)
Homogeneous Non-Newtonian Slurries
8 2
Bingham > τ0 R r Dτ0 4 1
> 1 - Rr - 2ζ 1 - 1- ζ þ ζ4
plastic < r ≥ r p 2η
B R 8ηB 3 3
>
: τ 0 R
p
>r<r ð1 - ζ Þ2
2ηB
8 " ! #
Casson > R
2
r 8 1 3
r 2
r
Dτ0 16 1 4ζ ζ4
>
> τ 0 1- ζ 2 þ -
model > 1- - ζ2 1 - þ 2ζ 1 - 8k 7 3 21
< r ≥ r p 2k R 3 R R
>
>
> τ R 8 1 3
>
: r < rp 0 1 - ζ2 - ζ 2 1 - ζ2 þ 2ζ ð1 - ζ Þ
2k 3
8 !
Herschel– 1 1þn 1þn 1
>
> nR τ0 n r n D τ0 n ζ 2 2ζ 1
Bulkley >
> n
ð1 - ζ Þ - -ζ ð1 - ζÞa þ ð1 - ζÞb þ ð1 - ζ Þc
< r ≥ r p ð 1 þ nÞ k R 2 k a b c
fluid 1 1 1
>
> 1 a=1 þ b=2 þ c=3 þ
> nR τ0 n 1þn n n n
>
: r < rp ð1 - ζÞ n
ð 1 þ nÞ k
165
166 6 Non-Newtonian Slurries and Suspensions
Table 6.3 Relationship between wall shear stress and viscosity as a function of radial location and
the viscosity at the wall in laminar flow for various non-Newtonian fluid types
Model η(r, τ0) η0
Newtonian μ μ
Ostwald–de Waele (power law) n - 1 n-1
τ0
r n τ0 n
R 1
1 kn
8 kn
Bingham plastic > ηB r ηB
< r ≥ rp
r 1-ζ
-ζ R
>
: R
r < rp 1
8
Casson model >
> r ≥ rp
k
!
r
η0 =
k
>
< 1 R 1
ζ-2 ζ
r 2 r
þ ζ - 2ζ 2 þ 1
>
> R R
>
:
r < rp 1
8 r τ0 τ0
Herschel–Bulkley fluid
< r ≥ rp
> 1 1 1
τ n r n τ0 n 1
R 0 -ζ ð1 - ζ Þn
>
: k R k
r < rp 1
( τy
τ0 > τy : D = ζD
Dp = 2r p = τ0 ð6:1Þ
τ0 ≤ τy : D
where:
τy
ζ= ð6:2Þ
τ0
To obtain the relationship between the wall shear stress and the bulk velocity (V )
for a fluid having a yield stress, the integration technique of Sect. 2.5 is performed
between the radial limits of rp–R rather than 0–R. Performing these integrations and
rearranging the terms allows the relationship between the wall shear stress and the
bulk velocity to be determined, while simple examination of the constitutive equa-
tions of the fluids provides the velocity profiles in laminar flow. Table 6.2 shows the
resulting equations for several non-Newtonian flow models commonly used for
industrial slurries. The variation of viscosity with radius is given by Eq. (6.3),
which has been applied to these same models in Table 6.3. Figure 6.4 provides
plots of the velocities and viscosities for these fluid models.
τ ðr Þ r τ0
η ðr Þ = duðr Þ
=- ð6:3Þ
R duðrÞ
dr dr
For shear-thinning fluids or slurries, the viscosity is lowest at the wall (where the
shear rate γ_ = - du=dr is a maximum). It is highest in the pipe center (or at the
boundary of the unsheared plug at the nondimensional radius (ζ), if the fluid has a
yield stress). It is worth considering what effect this has on particles that may be
6.2 Homogeneous Non-Newtonian Slurries 167
u(r)/Umax (-)
0.0 0.2 0.4 0.6 0.8 1.0 1.2
1.0
Newtonian
0.5
Power law U/Umax
r/R
0.0
-0.5
-1.0
0 0.5 1 1.5 2 2.5 3
h(r)/h0 (-)
u(r)/Umax (-)
0.0 0.2 0.4 0.6 0.8 1.0 1.2
1.0
0.0
-0.5
-1.0
0 5 10 15
h(r)/h0 (-)
Fig. 6.4 Radial variation of normalized velocity (upper half) and normalized viscosity (lower half)
in laminar pipe flows for fluids with and without a yield stress. (Note that for the model parameters
used for illustrative purposes in this text, the Casson and Herschel-Bulkley model fluid values are
essentially the same; thus, the Casson model has been omitted for clarity)
suspended in such fluid. Figure 6.5 shows the results of a simple simulation of a
single particle being conveyed through a pipe under laminar flow for these four fluid
types, where the viscosity at the wall (and hence the pipe Reynolds number) is the
same for each fluid. The particle is released from the top of the pipe on the plane of
symmetry.
The radial variation in viscosity dramatically slows the particle’s descent in the
central core of the pipe. Indeed, in this simulation, the variation prevents the particle
from entering or crossing the unsheared plug-in fluids having a yield stress. It will be
shown in the following sections that this effect influences the particles’ solid
distributions in heterogeneous non-Newtonian flows, which in turn modifies the
required transport pressure gradient.
168 6 Non-Newtonian Slurries and Suspensions
1
0.8
0.6
0.4
0.2
r/R (-)
0
-0.2 Newtonian
-0.4 Power law
z
-0.6 Bingham plasc
-0.8 Herschel-Bulkley model
-1
Axial distance
Fig. 6.5 Trajectories of a particle settling through the four fluids, where the wall viscosity and wall
shear stress are the same for each case
11:6μ
δ= ð6:4Þ
ρU
where μ and ρ are the viscosity and density, respectively, of the Newtonian fluid, and
pffiffiffiffiffiffiffiffiffi
U* is the shear velocity at the wall (i.e., τ0 =ρ). In the main flow, it is assumed that
all momentum transfer takes place by turbulent mixing, which is an inertial process
rather than a viscous one. This mixing gives rise to the logarithmic velocity profile in
the turbulent region (see Eq. (2.36) and the associated text), as opposed to the linear
velocity increase in the sublayer, where:
3
Comparing this equation with Eq. 2.36 we see that the layer thickness is here based on the
engineering approximation of y + = 11.6, rather than the actual value of y + ~5.
6.2 Homogeneous Non-Newtonian Slurries 169
τ0 y
u= y<δ ð6:5Þ
μ
As the velocity gradient in the viscous sublayer is much steeper than in the
logarithmic zone, the sublayer has an influence on mean velocity that is dispropor-
tionate to its small thickness (see Fig. 2.14). As seen in Sect. 2.5, it is the sublayer
that determines the influence of viscosity (via the Reynolds number) on the friction
factor for Newtonian turbulent flow.
If a non-Newtonian fluid is to be substituted for the Newtonian one, certain
reasonable predictions can be made. In the logarithmic zone, the momentum trans-
port is inertial in nature; as a result, the velocity gradient is not affected by
Newtonian viscosity, and it may be expected that the same would apply to
non-Newtonian rheological properties. (Due to a non-zero τy, some change in the
velocity profile would be expected in the part of the flow nearest the pipe axis;
however, the effect on mean velocity is very small.) Within the sublayer, which
occupies such a small portion of the flow that variations in shear stress and velocity
gradient are negligible, laminar conditions are approached. Again, the linear velocity
increase applicable to Newtonian fluids should also apply to a non-Newtonian fluid.
Two questions arise at this point. The first concerns the value of μ that applies in
Eqs. (6.4) and (6.5), which for non-Newtonian fluids, will be designated η from here
on. The second refers to the coefficient that specifies the thickness of the viscous
sublayer. (For Newtonian flows, this is 11.6, as given in Eq. (6.4), but the value may
be larger for non-Newtonian flows, as discussed below.)
We can resolve the first question by referring to Fig. 6.6. The “secant” slope on
the figure (η) is equivalent to the Newtonian viscosity, in that it represents the ratio of
τ to du/dy, but only for the shear rate at which this line intersects the rheogram. It is
not constant like the Newtonian viscosity, but must be expressed as a function of the
shear rate (du/dy) or the shear stress (τ). Thus, Eq. (6.5) should still apply for
non-Newtonian fluids, provided μ is understood as the general viscosity (η) of
Fig. 6.6 and evaluated at τ = τ0 (i.e., η0).
The remaining question, concerning thickening of the viscous sublayer, can be
related to a conceptual model proposed by Lumley (1973, 1978) to explain the
phenomenon of drag reduction in aqueous flows, obtained by the addition of small
typical homogeneous
mineral slurry
thickness of
d viscous sublayer
eddy size
quantities of certain long-chain molecules. These substances act to increase the size
of the smallest dissipative, turbulent eddies. Figure 6.7, which illustrates Lumley’s
model, graphs the distance from the wall ( y), as the ordinate, and the representative
eddy sizes as the abscissa. As indicated in the figure, the size of the large eddies (the
macro scale of turbulence) is directly proportional to the distance from the wall,
while the size of the smallest eddies (the dissipative, or Kolmogorov, scale) changes
little with this distance. At large values of y, the inertial macro-eddies are much
bigger than the dissipative micro-eddies, and between them is a whole range of
turbulent eddy sizes, represented by the shaded area in the figure. As the wall is
approached, the range of possible eddy sizes shrinks until the sizes of the largest and
smallest eddies are equal, indicating the elimination of turbulence. At this point,
y equals δ, the thickness (in a statistical sense) of the viscous sublayer.
The dashed line in Fig. 6.7 shows that an increase in the size of the dissipative
micro-eddies leads to an increase in the thickness of the viscous sublayer. It is
known that the micro-eddy size is increased by long-chain molecules and that
non-Newtonian rheological properties can produce a similar effect, as shown by
Wang and Larsen (1994). If other quantities are unaffected, the thickened sublayer
will, in turn, produce a higher mean velocity for the same wall shear stress, giving a
lower friction factor.
A predictive model for non-Newtonian fluids has been developed on this basis
(Thomas and Wilson 1987; Wilson and Thomas 1985). It was noted that the
interaction of eddies in turbulent flow causes them to increase in axial length,
often very rapidly. This “vortex stretching” process produces abrupt increases in
shear rates and in the viscous dissipation of the smallest eddies. As the energy
available for dissipation is fixed by the “turbulent energy cascade,” the result is an
increase in the micro-eddy size. The model uses the ratio of the integrals under the
non-Newtonian and equivalent Newtonian rheograms—denoted by α and defined
6.2 Homogeneous Non-Newtonian Slurries 171
previously4 in connection with Fig. 6.6—to estimate the size increase of the micro-
eddies. From the relationship shown in Fig. 6.6, it follows that the thickness of the
sublayer should also be multiplied by a factor equal to α. As shown by Wilson and
Thomas (1985), the value of the mean velocity (Vm), which results from this
thickened viscous sublayer, is given by
Vm VN
= þ 11:6ðα - 1Þ - 2:5 lnðαÞ ð6:6Þ
U U
where VN is the mean velocity for the equivalent smooth-wall flow of a Newtonian
fluid with viscosity μ0 = η0, and the deviation from the logarithmic profile near the
pipe axis has been ignored (see Eq. 2.37). Since the velocity profile in the viscous
sublayer is linear near the wall, the viscosity at the wall (η0) can be simply computed
from the fluid’s constitutive equation, recognizing that η0 = τ0 =_γ 0 . Wall viscosity
values for the fluids described in Table 2.3 are given in Table 6.4.
Considering the region within the pipe where the applied stress was less than the
yield stress, Wilson and Thomas (1985) also postulated that this region would have a
flatter turbulent velocity profile, analogous to the unsheared plug under laminar flow,
and modified Eq. (6.6) to account for this as follows:
Vm VN
= þ 11:6ðα - 1Þ - 2:5 lnðαÞ - Ω ð6:7Þ
U U
4
In Fig. 6.6, the equivalent Newtonian-fluid rheogram is the triangular area defined by the vertical
line indicating the shear rate of interest and the line defining the viscosity, and the integral
ANewtonian, is simply the area of this triangle. The non-Newtonian fluid rheogram is the full curve,
up to the same shear rate of interest, and the integral Anon-Newtonian, is simply the integral under this
curve. α = Anon-Newtonian/ANewtonian.
172 6 Non-Newtonian Slurries and Suspensions
Fig. 6.8 Comparison between a kaolin slurry pipeline data and predictions based on the analytical
laminar equation (Table 6.2) and Wilson and Thomas method (Eq. (6.6)) using a Herschel–Bulkley
model fluid. The Herschel–Bulkley model parameters are given, and the transition velocity, based
on a Bingham plastic model using Eq. (6.11), is shown as a dashed line. (Experimental data from
GIW Hydraulic Lab)
6.2 Homogeneous Non-Newtonian Slurries 173
Both Figs. 6.2 and 6.8 demonstrate that laminar and turbulent flows can be viable
solutions for pumping non-Newtonian slurries. Since the mechanisms and compu-
tational methods differ considerably, it is important to be able to predict whether a
flow might be laminar or turbulent. For Newtonian fluids, the Reynolds number
alone defines whether a flow is turbulent or not, and for industrial installations, a
value of 2100 > Re > 2300 can be used, although much higher values can be
obtained in laboratories where perturbations are reduced to a minimum.
For power-law fluids, Metzner and Reed (1955) were able to show through data
fitting that transition would occur at a critical friction factor, f = 0.0304. Their
conclusions were based on analysis using a generalized Reynolds number, ReMR,
resulting in a critical Reynolds number:
8ρf V 2
Re MRc = n ≈ 2100 ð6:8Þ
K 0 8V
D
8ρV 2ann
Re 3 = n ≈ 2100 ð6:10Þ
τy þ k 8V ann
Dshear
where the average velocity in the sheared annulus Vann = (V - Vpζ 2)/(1 - ζ 2).
Vp = u(0, τ0), which can be found in Table 6.2, and the sheared diameter is defined as
Dshear = D(1 - ζ).
Clearly, as might be expected, the ratio of the inertial forces to the viscous forces
(i.e., the Reynolds number) is important in describing the transition from laminar to
turbulent flow. However, over the years, many researchers have found that for fluids
with yield stresses, in particular Bingham plastics (e.g., Govier and Aziz 1972;
Slatter and Wasp 2006; Thomas and Wilson 2007; Wilson and Thomas 2006), a
174 6 Non-Newtonian Slurries and Suspensions
far simpler relationship for calculating the transition velocity (Vt) may be used,
namely:
rffiffiffiffiffiffiffi
τyB
V tB = kt ð6:11Þ
ρf
where 20 < kt < 26, depending on the database used. This relationship—shown
graphically in Fig. 6.9b—is independent of pipe diameter, and hence the pipe
Reynolds number. This rather surprising assertion is worth closer examination.
Plotted in Fig. 6.9a are the transition velocities obtained from the intersection of
the laminar curves of Table 6.2 and the turbulent curves obtained from Eq. (6.7) and
Table 6.4 for some typical Bingham plastic and Herschel–Bulkley fluids. These are
compared with those predicted by Eq. (6.11) for 20 < kt < 26 (shaded area).
It can be seen that for chosen fluids and for industrially relevant pipe sizes, i.e.,
D ≥ 100 mm, predictions based on this simple equation are good and provide a
simple means of calculating the transition velocity. Case Study 6.1 provides an
example of how to use this method for fluids other than Bingham plastics.
Apart from flow regime delineation, there is another important reason to examine
transitional flow of homogenous slurries: that of optimization. The rheology of
typical non-Newtonian slurries and their resulting transport pressure gradients are
strong functions of solids concentration. Both contribute to determine the specific
energy consumption (SEC), which is proportional to both greenhouse gas equivalent
generation and operating costs (OPEX). Consider the transport characteristics of
some brown coal/water mixtures shown in Fig. 6.10.
Figure 6.10 is typical of the behavior of homogeneous slurries, which is readily
explained if one considers the requirements for a given throughput of solids.
Consider a slurry that is initially conveyed at a relatively low concentration and
high velocity in turbulent flow. Increasing the solids concentration has a small effect
on the rheology, but directly lowers the transport velocity, as well as the associated
transport pressure gradient required to deliver a given tonnage, thus reducing the
SEC. As concentration increases and velocity decreases to maintain a fixed tonnage,
the transition from turbulent to laminar flow is reached. As shown in Fig. 6.2, the
change in pressure drop in laminar flow is now a weaker function of velocity but a
strong function of solids concentration. Increasing the solids concentration further
increases the required pressure gradient at such a rate that SEC again increases,
producing a minimum at or near the transition velocity.
Transitional flow is characterized by intermittent turbulent flow, where turbulent
“puffs” are separated by periods of laminar flow. If there is a fear that the slurry may
not be entirely homogeneous, such actions gently and periodically provide mixing in
the pipe that may be used to resuspend any particles that have settled. While this
condition would appear optimal for transporting homogeneous slurries, there are
some further considerations when transport is provided by centrifugal pumps. These
will be addressed in Chap. 12.
6.2 Homogeneous Non-Newtonian Slurries 175
5.0
20
3.5
0
3.0 0 500
Shear rate (1/s)
2.5
2.0
1.5 (a)
1.0
0.5 (a)
0.0
0 200 400 600 800 1000
Pipe diameter (mm)
9
8 Sm
Sm == 1.1
1.1
Sm == 1.3
Sm 1.3
7 Sm
Sm == 1.5
(m/s)
1.5
velocity(m/s)
6 Sm
Sm == 1.7
1.7
Sm = 1.9
Sm = 1.9
Transition velocity
5
4 (a)
Transition
3 industrially relevant
velocities
2
1 (b)
0
0 20 40 60 80 100
Yield stress (Pa)
Fig. 6.9 Effect of pipe diameter and rheological properties on transition velocity. (a) Comparison
of the transition velocities obtained from the intersection of the laminar and turbulent curves
produced by the Wilson and Thomas (1985) method for two different rheological models (shown
inset), compared with those predicted by Eq. (6.11) for the range of values of kt (shown shaded). (b)
Variation of the transition velocity with suspension yield stress and suspension relative density (Sm),
obtained from Eq. (6.11) with kt = 26. (Pullum et al. 2018)
176 6 Non-Newtonian Slurries and Suspensions
In the early 1970s, pumping tests were being conducted to assess the attrition of
coarse coal particles using conventional hydraulic transport methods (Elliot and
Gliddon 1970). After a period of pumping, it was discovered that the mode of
transport had changed: the pressure gradient no longer exhibited the characteristic
hook and associated minimum conveying velocity of turbulent coarse particle
(settling slurry) transport (Sect. 3.4), but rather showed a linear relationship. Fur-
thermore, transport (lack of settling) continued down to very low velocities with
surprisingly high solids concentrations. The characteristics appeared to be those of
Bingham plastics. In addition, tests conducted with similar suspensions at different
pH levels required dramatically different transport pressure gradients. All the data
implied that the suspensions were behaving as homogeneous non-Newtonian slur-
ries, but the particle sizes were in the range of 12.5 mm—two orders of magnitude
greater than the particle sizes discussed in Sect. 6.1.1.
Before going further, it is worthwhile to look more closely at this claim that only
particles less than, say, 50 μm will combine with the conveying liquid to form a
non-Newtonian slurry or carrier fluid in which all larger particles are conveyed. The
exact values of the various inter-particle forces are difficult to calculate, as they are
dependent upon a wide range of properties that affect the electro-chemical environ-
ment in which the particles exist. Because of this, even if these calculations can be
made, they cannot be generalized from one slurry to another and so it becomes an
arduous and futile exercise. Similarly, experiments will only produce quantitative
6.3 Heterogeneous Non-Newtonian Slurries 177
5
Colloidal science would claim a much smaller top size, of order 1–2 μm, which are typical of clay
particles; however, 50 μm provides a pragmatic practical limit, achievable in laboratories and
on-site.
178 6 Non-Newtonian Slurries and Suspensions
Fig. 6.11 Variation of yield stress for (a) three different thickened suspensions and for (b) the same
three suspensions but using only the <38 μm fractions. (Coghill et al. 2014)
6.3 Heterogeneous Non-Newtonian Slurries 179
Fig. 6.12 Transport characteristics of Stab-flo mine waste suspensions. (Original presentation by
Duckworth et al. 1986)
yield stress of the carrier fluid exceeded that required to support the particle (i.e.,
τy > τyc Eq. (2.68)), then the coarse solids would be lifted away from the pipe wall
through the action of shear and rotation (e.g., Saffman and Magnus forces at the pipe
wall). They would then be held in the unsheared core of the pipe where they would
not interact with the pipe wall and would only increase the conveying gradient by
virtue of increasing the carrier fluid density, and possibly through modification of the
extent of the central unsheared core. This concept appeared to work, as demonstrated
by these data in Fig. 6.12, obtained at the GIW Hydraulic Laboratory.
The data apparently show behavior in both the turbulent and laminar regimes, and
the curves shown are predictions based on the carrier rheology (lowest curve) and
coarse particle concentrations. Scale-up was successfully achieved using these
techniques, although only over a relatively small pipeline range, DN150 to
DN300. The suspensions were characterized by broad particle size distributions
and the ability to be stopped and restarted at will. Closer observation of the flow
regimes was required to ascertain whether this boosted pseudo-rheology was real.
180 6 Non-Newtonian Slurries and Suspensions
With the aid of magnetic resonance imaging (MRI), Pullum and Graham (1999,
2000) demonstrated that, as predicted by Thomas, coarse particles suspended in a
fluid with a yield stress five times greater than required to support the particles (i.e.,
τy > 5τyc or ξ = τy/τyc = 5), migrated across a horizontal pipe to form a sliding bed
of solids (Fig. 6.13). Other practitioners observed similar behavior in the field
(Cooke 2002). This was in contradiction to the mostly uniform concentration profiles
that had been observed in earlier research, although it should be noted, in defense of
the earlier work, that the profiles had been obtained at a very high total solids
concentration where the particle mobility was low and the contrast between the
carrier fluid and coarse burden was very low. Clearly, under laminar flow, most
broadly sized distributions do not behave as a homogenous slurry.
The broadly sized distribution of Stab-flo suspensions makes it difficult to
differentiate between the actions of the underlying carrier fluid slurry and those of
the coarse burden it is conveying. To overcome this problem, the carrier fluid can be
modeled using a clear aqueous polymer gel with similar rheology to the mineral
slurry (e.g., Carbopol), and coarse particles that do not contain the middle solids
fractions (say, 50–200 μm). These artificial analog suspensions model the main
features of interest in the real suspensions (i.e., the non-Newtonian fluid dynamics
and the motion of the large particles), while their transparency allows visual obser-
vation. Comparison of the measurements shown in Fig. 6.13 for such analog
suspensions with those shown in Fig. 6.14 (a broadly sized distribution of a real-
world suspension) demonstrates that the analog suspension appears to capture the
main features of a “real” suspension.
Section 2.7 noted that for a suspension where ξ > 1, particles would not settle
unless the fluid was sheared. Furthermore, Fig. 6.3 shows that in a homogeneous
pipe flow, there will be an unsheared plug with a radius ζ R, yet in the concentration
profiles shown in Figs. 6.13 and 6.14, the solids have settled across the entire pipe
cross section to form a bed. This occurs because the axisymmetric stress profiles
shown in Fig. 6.3 become distorted as solids that are free to settle in the sheared
annulus start to form a bed. Irrespective of whether this bed is stationary, sheared, or
unsheared, the bed’s cross section restricts the fluid flow through the lower invert of
the pipe, modifying the flow from axisymmetric to plane-symmetric (Fig. 6.15).
The result, depicted in Fig. 6.15, is that the unsheared plug, which can support the
particles, reduces in cross section and its lower contour moves upwards. In doing so,
it exposes particles that were previously contained within the lower portion of the
concentric plug to shear, allowing them to settle. Depending on the depth of the bed
and the carrier stress ratio (ζ), the newly positioned, unsheared region may sweep
through the entire original plug back into areas of the pipe that were previously
sheared. Under these circumstances, solids in the original concentric plug will be
subjected to shear and will settle, removing all coarse particles from the carrier fluid
above the bed, and forming a completely stratified flow (Fig. 6.13). If, on the other
hand, the starting and ending plug locations overlap, a remnant of the original
6.3 Heterogeneous Non-Newtonian Slurries 181
concentric core will remain filled with solids at the original feed concentration to
form a smaller plug traveling above the bed through the surrounding fine-particle
carrier fluid (Fig. 6.16). Such flows, where the coarse fraction of the solids are
Fig. 6.13 MRI concentration and axial velocity maps and profiles for <2 mm sand in Carbopol. (a)
At incipient stationary bed formation. Velocity and concentration profiles are drawn through the
vertical diameter, and across a chord passing through the dynamic center. Concentration profiles are
arbitrarily scaled. (b) Bed development during transport showing concentration maps (black/white)
and velocity maps (color). (Pullum and Graham 2000)
182 6 Non-Newtonian Slurries and Suspensions
mobile within the carrier fluid, are known as complex suspension flows. Similar
behavior has also been observed in channel flows (Spelay 2007).
Since the solids in laminar flow settle to form either a sheared or an unsheared
bed, and since suspension, if any, is very limited, a natural model for these flows is a
stratified—or two-layer—model similar to that described in Chaps. 4 and 5. Such
models are more complex, since the various interfacial stresses and flow beneath the
bed must account for the spatial variation in shear stress, and hence the viscosity of
the carrier fluids (Fraser and Goosen 2019; Pullum et al. 2004; Rojas and Saez
2012). Examining the concentration and velocity profiles in Fig. 6.13a, where the
solids are totally contained in the bed and the velocities of the bed and flow above the
bed may be considered as constant values with reasonable accuracy, one can see that
these profiles replicate the simple nature of a two-layer model. As before, the models
use fluid properties (in this case, those of the carrier fluid slurry), the density and size
distribution of the coarse burden, the degree to which these solids will pack in the
bed, and some form of frictional coefficient depending on whether or not the friction
may be considered Coulombic (Talmon et al. 2014). These latter properties are not
immediately obvious, since the carrier fluid’s high viscosity affects the neighboring
particles’ proximity to each other, and to the bounding surfaces, through a squeeze
film effect, so that volumetric bed concentrations are typically 5–10% lower than
they would be in water. Results of calculations obtained with such a model are
shown in Fig. 6.17, and predictions for inclined flows using this model are compared
with laboratory data in Fig. 6.18.
6.3 Heterogeneous Non-Newtonian Slurries 183
Fig. 6.14 Sheared settling of a backfill paste in a 100 mm pipe as shown from Electrical Resistance
Tomography concentration maps. (Graham et al. 2017)
Fig. 6.15 Variation of fluid shear stress with bed depth. (a) Schematic showing the fluid’s linear
stress distribution on the vertical axial plane with and without a stationary bed of solids. (b)
Computed stress distributions for a visco-plastic fluid flowing through a series of truncated
cylindrical channels with hydraulic diameter (Dh), at a constant flow rate and resulting decreased
stress ratio (ζ). The yield stress contour, delimiting the unsheared plug, is approximated by the
yellow ellipses with major and minor axes ( p and q, respectively). (Pullum et al. 2018)
184 6 Non-Newtonian Slurries and Suspensions
Fig. 6.16 Plug flow above a sliding bed. Concentration profiles obtained from electrical resistance
tomography (ERT) imaging of a Carbopol/glass analog suspension
37% v/v
3.0
43% v/v
2.5
2.0
1.5
1.0
0.5 (b)
0.0
0.0 1.0 2.0 3.0
Velocity (m/s)
6.3 Heterogeneous Non-Newtonian Slurries 185
7.0
+20°
6.0
5.0
+12°
Pressure gradient (kPa/m)
4.0
+5°
3.0
+0°
2.0
-5°
1.0
0.0
-12°
0.0 0.5 1.0 1.5 2.0 2.5
-1.0
-20°
-2.0
-3.0
Velocity (m/s)
Fig. 6.18 Transport characteristics for various pipe inclinations. (Pullum et al. 2006)
6
Similar behavior had been observed for fully stratified flow in water based coarse particle
suspensions producing transport pressure gradient of the same order (Babcock 1971; Newitt et al.
1955).
186 6 Non-Newtonian Slurries and Suspensions
10 2.5
tailings, rs =2650 kg/m3, cv = 30% v/v -40mm ROM
coal
1.0 1 kPa/m
0.1
industrially Homogenous
relevant 0.5
laboratory
scale Constant delivered tonnage
0.0
0.01
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 100 200 300 400 500
Pipe diameter (m) Pipe diameter (mm)
(a) (b)
Fig. 6.19 Variation of pressure gradient with pipe diameter: (a) comparison of pressure gradient
predictions for homogeneous and fully stratified suspension models; (b) ultra-high concentration
run of mine (ROM) coal. (Pullum et al. 1996)
tests are performed using small pipes (say, D < 100 mm), which can lead to full-
scale designs that grossly under-predict the full-scale pressure gradients, as shown.
Figure 6.19b compares a two-layer model prediction with data obtained from
large-scale pipelines conveying high concentrations of ROM coal, a material with a
very broadly sized distribution, the finer fractions of which combine with the water
to form a non-Newtonian carrier. Similar behavior was also observed when pumping
broadly sized distributions of mine waste. A common rule of thumb in the pipeline
industry is that stratified flows will require a transport pressure gradient of 1.5–2 kPa/
m, which closely approximates the stratified model predictions in Fig. 6.19a. This is,
however, a 50–100% over-prediction for the coal suspension in Fig. 6.19b, even
though the volume concentration is approximately double. The variation in the
transport gradient is primarily due to the difference in submerged densities of the
particles, but it is also due to the mobility of the particles which is a function of the
grain size, concentration, and flow regime.
7
While turbulent, generally these flows are at much lower pipe Reynolds numbers than for water-
based suspension, typically 103 cf. 105.
188 6 Non-Newtonian Slurries and Suspensions
Fig. 6.20 Typical turbulent viscosities in pipeline flow of a visco-plastic fluid flowing at 2.9 m/s
through a DN50 pipe. (a) Snapshot of the normalized instantaneous viscosities across the pipe. (b)
Variation of the normalized viscosity across the pipe. ReG is the wall Reynolds number defined
using the viscosity at the wall (η0), defined in Table 6.4. (Singh et al. 2016)
cvd = 30%
equivalent Newtonian
model
cvd = 17%
non-Newtonian
model
Fig. 6.21 Chord-averaged ERT concentration profile across a pipe showing the difference in
predicted profiles based on a Newtonian fluid and a non-Newtonian fluid each with the same
viscosity at the pipe wall. (Matoušek et al. 2021)
suspension above that which was found for equivalent Newtonian flows, retaining
the particles in this more viscous core, as seen in Fig. 6.21. In this figure, both the
Newtonian viscosity and the non-Newtonian wall viscosity are the same (i.e., the
6.3 Heterogeneous Non-Newtonian Slurries 189
Fig. 6.22 Non-Newtonian suspension predictions: (a) sand in Carbopol showing the underlying
carrier fluid, laminar flow modeled with a non-Newtonian two-layer model and turbulent flow
modeled with a modified multi-component model. (Pullum et al. 2015); (b) thickened uranium
tailings modeled with a modified multi-component model. (Coghill et al. 2014)
pipe Reynolds numbers are the same). The non-Newtonian version captures the
enhanced solids suspension much better and corroborates the notion that solids are
more readily suspended in non-Newtonian flows. This, combined with the higher
viscosities, means that sliding beds or hugely stratified flows are unlikely to occur for
small- to medium-sized particles in turbulent flow—although very large or massive
particles will still form sliding beds.
The ability to predict the settling and suspending behavior of solids in sheared
non-Newtonian flows (provided that a local viscosity calculated from the local shear
rate is used) is encouraging. It suggests that Newtonian heterogeneous hydraulic
transport models might be applied to non-Newtonian systems, as long as similar
precautions are taken and the spatial variation of the instantaneous viscosity is
employed. Results obtained using a simplified 4CM (Sect. 5.5.1) are shown in
Fig. 6.22.
In this three-component model, the fluid and pseudo-homogeneous fractions are
combined into one: the carrier fluid slurry. The model is still under development and
is, at the time of this writing, awaiting rigorous testing and analysis at the GIW
Hydraulic Laboratory. To assess whether the model could be used with minimal
development, the three contributory fractions (i.e., the carrier fluid and the hetero-
geneous and stratified fractions) have simple multipliers—k1, k2, and k3,
respectively—associated with them. Fits between the data and the model are then
achieved by manipulating these values. A parity plot of the data so far (Fig. 6.23)
shows that considerable refinement will be needed beyond simple replacement of the
fluid and homogeneous fractions by the carrier fluid slurry properties.
This is not surprising, since it is within the carrier fluid that the non-Newtonian
rheology has the greatest effect (e.g., Fig. 6.21), while the coarser fractions appear
unaffected by the change in rheology, provided the appropriate viscosity is used.
Although the model is in early development, these figures show the potential for
190 6 Non-Newtonian Slurries and Suspensions
Fig. 6.23 Parity plot of observed data versus three-component model predictions. Closed
symbols—sand/Carbopol suspensions; open circles—glass/Carbopol suspensions; X—uranium
tailings. (Pullum et al. 2015)
This century has seen the development of highly thickened, high solids concentra-
tion pastes for use in tailings disposal via central thickened discharge systems and
back-fill operations. Although this form of disposal has not been adopted to the
extent of the aforementioned thickened tailings, it is a burgeoning development area.
Due to the high viscosity of these systems, most flows appear to be laminar.
However, as demonstrated in Fig. 6.14, even suspensions with static yield stresses
in excess of 400 Pa will stratify when sheared. Depending on the solids’ size
distribution and the rheological activity of the fine particles, these suspensions
should be considered to behave as either homogeneous or (more likely) heteroge-
neous suspensions.
The basic process for above-ground operations involves the concentrating and
thickening of a mineral plant waste stream via a thickener (high-rate or paste-type),
normally using flocculating agents. The underflow from these thickeners is highly
6.4 Conversion and Scale-Up 191
non-Newtonian. Sands may be added at this stage, or else downstream of the pumps,
which are now fed with this underflow. These pumps may be purely centrifugal or a
combination of centrifugal and positive displacement. The highly viscous nature of
these slurries means that pump performance is derated, as described in Chap. 9. The
flow, from both the thickener to the pump and from the pump to the tailings storage
facility (TSF), is non-Newtonian, requiring the analysis given in this chapter.
Upon discharge from the end of the pipeline, the slurry spreads across the TSF via
sheet flow or channel flow, both non-Newtonian. The exact nature and extent of the
flow on the TSF depends primarily on the discharging slurry’s yield stress. The flow
may be confined to a conical deposit radiating from the discharge points, or it may
flow to form a lower-graded deposit. The high concentration of such deposits
removes the dangers associated with conventional TSFs, where the overpressure
of the excess water and the mobility of the particles can combine to produce
catastrophic failure of the tailings dam. Interested readers are directed to the hand-
book Paste and Thickened Tailings: A Guide (Jewell and Fourie 2015) for more
information on this area of hydraulic conveying.
The computational power that is readily available to engineers these days means that
scaling up of non-settling behavior is trivial if the slurry or fluid can be described by
a suitable rheological model, and it is only marginally more difficult if the raw data
alone are used. Suitable techniques have been described in the previous sections and
the following case studies provide examples of how this may be achieved. However,
a powerful technique developed by Rabinowitsch (1929) and Mooney (1931)
enables laminar flow of a homogeneous fluid, regardless of rheological type, to be
scaled up using appropriate affinity laws, similar to those used in pump design. This
technique is included here for completeness.
6.4.1 Conversion
Section 2.5 showed that the stress in the fluid at the wall (τ0) is a simple function of
the pressure gradient and pipe diameter, regardless of fluid type (Eq. (2.25)). It also
showed that for a Newtonian fluid, the corresponding shear rate at the wall (i.e., the
velocity gradient at the wall) is γ_ 0 = 8V=D. For non-Newtonian fluids, Rabinowitsch
and Mooney established that the shear rate at the wall is generally given by
ð3n0 þ 1Þ 8V
γ_ 0 = ð6:12Þ
4n0 D
where:
192 6 Non-Newtonian Slurries and Suspensions
Fig. 6.24 Conversion of pseudo shear rate data to true shear rate data using the Rabinowitsch–
Mooney relationships
d lnðτ0 Þ
n0 = ð6:13Þ
d lnð8V=DÞ
Note that for a power-law fluid, n′ n, but that n′ is generally a weak function of
8V/D. Since the actual wall shear rate is proportional to 8V/D, this value is denoted
the pseudo shear rate, and plots of τ0 versus 8V/D are known as pseudo shear
diagrams or pseudo-rheograms. Pseudo shear diagrams can be converted into true
shear stress/shear rate rheograms and vise versa by the application of Eqs. (6.12) and
(6.13), as shown in Fig. 6.24. This conversion enables the true fluid properties to be
used for designing other processes (such as mixing).
The laminar flow of any given fluid can be scaled from one pipe size to another using
the Rabinowitsch–Mooney technique. The only stipulations are that there is no “slip
of the fluid” at the wall (see Metzner 1961) and that the flow remains laminar in both
pipes. (Turbulent flow will be dealt with separately in the following section.)
As mentioned in the previous section, the basis of the Rabinowitsch–Mooney
technique is the plot of wall shear stress (τ0) versus 8V/D. Rabinowitsch (1929) and
Mooney (1931) proved that for all steady uniform laminar flows in a pipe, τ0 and
8V/D are linked in a functional relationship. In other words, for a given slurry, the
values of both τ0 and 8V/D can be determined from experiments in a single pipe.
This experimentally determined plot, the pseudo shear diagram or function linking
the two variables, can then be applied to all laminar flows of the slurry in question,
whatever the pipe diameter.
For some other pipe, say of diameter D2, we can obtain the pressure gradient
dp/dx from the plotted values of τ0 by multiplying them by (4/D2). Similarly, the
values of the mean velocity in the pipe of diameter D2 equal the values of 8V/D
multiplied by (D2/8). Since the factors by which each of the coordinates of the
pseudo shear diagram are multiplied depend only on D2, a simple re-scaling of the
axes of this figure gives a plot of dp/dx versus V for a pipe of diameter D2. Scale-up is
6.4 Conversion and Scale-Up 193
thus achieved by transforming the data from the experimental pipe (diameter D1) to
the prototype (diameter D2) on a point-by-point basis. This process is similar to
scaling head and discharge data for a centrifugal pump by using the affinity laws (see
Chap. 8). For the pipe, the analogous quantities are the frictional gradient ( jm) and
the mean velocity (V ). With subscripts 1 and 2 denoting experimental and prototype
conditions, respectively, the scaling relations (affinity laws) for laminar flows in
pipes are thus:
D1
ðjm Þ2 = ðjm Þ1 ð6:14Þ
D2
and:
D2
V2 = V1 ð6:15Þ
D1
Since we are scaling the same fluid in different pipes, the essentially constant terms
can be separated out thus:
ρm U
V = 2:5U lnðDÞ þ 2:5U ln ð6:16Þ
η0
As for laminar flow, the combined use of Eqs. (6.14) and (6.17) allows scale-up of
data for turbulent non-Newtonian flow from one pipe size to another. The velocity-
scaling relation (Eq. (6.17)) was originally derived on the basis of a hydraulically
smooth boundary (Wilson 1986) for which Eq. (2.37) applies. However, it can be
shown that this relationship still holds for pipes of similar roughness.
It should be noted that neither the laminar nor the turbulent scaling relationships
will predict the transition velocity from one regime to the other. A scaled data point
might represent laminar conditions in one pipe size and turbulent conditions in
another, and the scaling for that point will, therefore, be invalid. This problem can
be addressed by obtaining and scaling both the laminar and turbulent data together,
using the appropriate method for each, and taking the maximum resulting value for
the pressure gradient (or shear stress) at each velocity. The new intersection point of
the separately scaled laminar and turbulent curves then represents the new (scaled)
transition velocity.
The use of computational fluid dynamic (CFD) methods to describe the flow of
non-Newtonian fluids undergoing laminar flow has become routine, provided the
fluids are inelastic and the flow geometries are not extreme. Such an approach is
unnecessary for normal pipeline flows of homogeneous fluids, since these can be
solved analytically (Sect. 6.3.1), but CFD tools do have uses in the design and
understanding of pipe fittings, valves, and junctions. Note, however, that when yield
stress fluids are considered, bi-viscous models must be used to avoid numerical
problems at low shear rates (i.e., η = minðη0 , ηðγ_ ÞÞ, where η0 is a limiting viscosity
to be used at or below a predetermined low shear rate).
To date, the authors are unaware of any well-established general turbulence
models, such as the familiar k–ε model used for high Reynolds number Newtonian
flows, that may be applied directly for RANS8 of turbulent flows of homogeneous
non-Newtonian fluids. One reason for this is that turbulent non-Newtonian flows,
when they are achieved in pump and pipeline flows, are at relatively low Reynolds
numbers—on the order of 2000–3000—where few Newtonian turbulence models
have been developed, and only very limited validation has occurred for some
specific cases. The low Reynolds numbers have, however, allowed for very detailed
DNS and LES studies to be made examining the nature of the turbulent flow in pipes
and channels. (See Sect. 6.4.2 for an example.) Since model non-Newtonian fluids
can be devised that are completely inelastic, such techniques can be used to test the
validity of established turbulent flow methods such as the Metzner-Reed method
8
A brief description of some CFD methods (e.g., RANS, LES, and DNS) can be found in Sect. 5.7.
6.6 Case Studies 195
(Rudman et al. 2004) and the Wilson–Thomas method (Singh et al. 2016). These
detailed studies have already provided the basis for developing comparable
non-Newtonian turbulence models suitable for RANS modeling (Gavrilov and
Rudyak 2016) and hopefully will drive the development of these models in the
near future.
For heterogeneous suspensions, multi-phase models such as the drift-flux model
have been used with some success in both pipeline and sheet flows across tailings
dams. At present, however, these techniques only provide qualitative results. Further
progress may result from studies using discrete element modeling (DEM) combined
with DNS flows that are under development.
To use Eq. (6.11), the slurry should be classified as a Bingham plastic; however,
many slurries depart from Bingham plastic behavior at lower shear rates, and indeed,
sometimes only raw pipeline data are available for analysis. Consider the data
previously presented in Fig. 6.8.
A simple way to obtain an estimation of the Bingham plastic yield stress is to
use a linear extension of the higher shear rate laminar data, as shown in Fig. 6.25,
noting the point of intersection with the vertical pressure gradient axis (in this
12
Kaolin slurry, rf = 1200 (kg/m3), D = 0.077 (m)
10
Pressure gradient (kPa/m)
vtB
6
4.38
4
0
0 2 4 6 8 10 12
Velocity (m/s)
Fig. 6.25 Data from Fig. 6.8 with Bingham approximation and resulting transition velocity
estimates
196 6 Non-Newtonian Slurries and Suspensions
case 4.38 kPa/m). The wall shear stress is obtained from Eq. (2.15), i.e., (τ0 = ΔpD/
4L = 4380 0.077/4 = 84.3Pa). Looking at the laminar curve equation for a
Bingham plastic in Table 6.2, it is evident that the last fourth-order term only affects
values at a low shear rate. As a result, we can approximate the Bingham model slurry
using the high shear rate data only by omitting this term and turning the equation into
a simple linear form. The resulting equation can be rearranged in terms of the wall
shear stress, as follows:
Dτ0 4
V= 1- ζ
8ηB 3
8V 4
τ0 = ηB þ τ ð6:18Þ
D 3 yB
From here, we can see that the Bingham yield stress is three-fourths of the
intercept (i.e., τyB = 0.7584.3 = 63.2 Pa). Using this value in Eq. (6.11), we can
calculate the range of transition velocities corresponding to the values of kt, shown in
Fig. 6.25.
Table 6.5 Laboratory data at slurry density ρm = 1230 (kg/m3), pipe diameter D = 77 (mm)
V (m/s) 0.00 0.15 0.33 0.67 1.00 1.33 1.67 2.00 2.33 2.67 3.00
Δp/L (kPa/m) 0.66 1.81 2.39 3.34 4.14 4.55 5.22 5.91 6.04 7.32 7.00
6.6 Case Studies 197
ΔpD
τ0 = ð6:19Þ
4L
ð3n0 þ 1Þ 8V
γ_ = ð6:20Þ
4n0 D
160
D = 0.077 (m)
140
120
100
Δ pD/4L (Pa)
80
60
40
20
0
0 50 100 150 200 250 300 350
8V/D (1/s)
From this pseudo-shear rate diagram, it is evident that the fluid appears to have a
yield stress and could be modeled as a Bingham plastic or Herschel–Bulkley-type
slurry. It is not necessary to classify the slurry type, however, before being able to
use this diagram. The data have been converted from a pressure gradient and
velocity, originally specific to the 77-mm pipe, to general shear stresses and shear
rates. Consequently, we can use this diagram directly to obtain the desired informa-
tion for the full-sized pipe. The target pipe diameter is 300 mm, and the transport
velocity is 1.5 m/s; therefore, 8V/D_target = 8 1.5/0.3 = 40.0 (1/s). By drawing a
line for this value to intersect the data, as shown by the red lines in the figure, the
corresponding wall shear stress for the target pipeline is found as τo_target = 48.5 Pa.
Using Eq. (6.19), we can see that the required pressure gradient in the target pipeline
is then found to be Δp/L_target = 0.65 (kPa/m).
Provided the target flow is laminar and only a few calculations are required, this
graphical technique provides a very simple and rapid solution. However, it is often
desirable to express the data as a non-Newtonian model, in order that it can be
readily used in computation for both pipeline calculations and other purposes (e.g.,
in mixing tank design). Fortunately, non-linear fitting procedures are available,
making it a simple task to convert such data into a rheological model.
For illustration, we will fit a Herschel–Bulkley model slurry to the data using
Microsoft Excel’s Solver Add-in, although a similar analysis could be performed
using many other applications. The analytical function relating wall shear stress to
velocity for laminar flow is found in Table 6.2, i.e.
1
D τ0 n ζ 2 2ζ 1
V= ð1 - ζ Þa þ ð1 - ζ Þb þ ð1 - ζ Þc ð6:21Þ
2 k a b c
1 1 1
a=1 þ b=2 þ c=3 þ
n n n
The spreadsheet below is set out to include column F containing a vector of estimates
of the Herschel–Bulkley parameters τy, k, and n, the calculated values of V and V′
using these parameters, and the wall shear stress data. Column G contains the square
of the error between the actual velocity (V ) and V′, which is summed in cell G6,
ignoring the zero-velocity row. The Solver Add-in is then activated to minimize this
sum by modifying the estimates of τy, k, and n, the results of which are shown in
Fig. 6.27. The resulting curve fit is also shown graphically in Fig. 6.28.
Using these model parameters, the pressure gradient at 1.5 m/s in the target
pipeline is calculated to be 0.65 (kPa/m), which agrees with the value found
previously by direct scaling from the data.
Having a model for the fluid enables far more calculations to be made; for
example, we can calculate turbulent flow using Eq. (6.6), with a value for α found
in Table 6.4. We can then calculate the flow curve for both laminar and turbulent
flow using the maximum values of Eqs. (6.6) and (6.21), as shown in Fig. 6.29.
6.6 Case Studies 199
Objecve
=$B$2/2*(D10/$F$3)^(1/$F$4)*((E10^2)/$H$2*(1-
E10)^$H$2+2*E10/$H$3*(1-E10)^$H$3+1/$H$4*(1-E10)^$H$4)
8
D = 0.077 (m)
7
5
Δ p/L (kPa/m)
4 2 1
= 1− + 1− + 1−
2
3
2 ty = 11.7 (Pa)
k = 3.05 (Pasn)
1 n = 0.63 (-)
0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
V (m/s)
2.5
D = 0.3 (m)
Pressure gradient (kPa/m)
2.0
1.5
1.0
0.5
0.0
0 1 2 3 4 5 6
Velocity (m/s)
Fig. 6.29 Flow curve for both laminar and turbulent flow
250
30% w/w
150 20% w/w
100
50
0
0 100 200 300 400 500
shear rate (1/s)
Using the relationships shown in Fig. 6.31, the wall shear stress at any concen-
tration or shear rate can then be calculated with
Δp
SEC =
3:6cv Ss ΔL
Δp/ΔL is in kPa/m and can be obtained from the maximum values of Eqs. (6.6) and
(6.21), and noting the relationships:
1 ρm c w _ s = Qcv ρs , πD2
ρm = , cv = , M Q= V
cw
þ 1 -ρ cw ρs 4
ρs f
1
ρm = = 1111:9 kg=m3
0:15
3040 þ 1- 0:15
1000
1111:9 0:15
cv = = 0:055
3040
τy = e18:840:15-3:655 = 0:44Pa
202 6 Non-Newtonian Slurries and Suspensions
k = e14:780:15-3:86 = 0:19Pa:sn
236:3
Q= = 1:413m3 =s
0:055 3040
4 1:413
V= = 9:8m=s
π 0:4292
Solving for laminar flow using Eq. (6.21), which must be solved iteratively for given
values of wall shear stress, τ0_lam is
1 1
a=1 þ = 2:304, b = 2 þ = 3:304
0:767 0:767
1 0:44
c=3 þ = 4:304, ζ =
0:767 τ0
1
0:429 11 0:767 0:042 2 0:04 1
V= ð1 - 0:04Þ2:304 þ ð1 - 0:04Þ3:304 þ ð1 - 0:04Þ4:304
2 0:19 2:304 3:304 4:304
= 9:23m=s
After iteration, it is found that τ0_lam = 11.48 Pa produces a velocity of 9.8 m/s.
Therefore, if the flow is laminar (which is unlikely for such a low concentration),
then the pressure gradient is obtained from Eq. (2.16):
ΔP 4 11:48
= = 107Pa=m
ΔL 0:429
For turbulent flow, we must use Eq. (6.6), which also must be solved iteratively
for appropriate wall shear stress values. Guessing a higher value than the laminar
case, say τ0_turb = 150 Pa, ζ = 0.0029, the friction velocity is then:
rffiffiffiffiffiffiffiffiffiffiffiffiffiffi
150
U = = 0:367m=s
1111:9
while from Table 6.4, the viscosity at the wall and rheogram area ratio are,
respectively:
6.6 Case Studies 203
1
0:190:767
η0 = 1 = 0:025Pa:s
ð150 - 0:44Þ0:767
0:767
α = 2 1- ð1- 0:0029Þ = 1:134
1 þ 0:767
After iteration, it is found that the wall stress τ0_turb = 188.8 Pa, and so:
ΔP 4 188:8
= = 1760Pa=m
ΔL 0:429
Considering Fig. 6.8, it is clear that this higher pressure gradient for the turbulent
flow regime is the correct value to use. Indeed, in general, the wall shear stress value
to use is simply τ0 = max (τ0_lam, τ0_turb).
The specific energy consumption (SEC) is then (note pressure gradient in kPa/m):
1:76
SEC = =2:93kWh=ðtonne:kmÞ
3:6 0:055 3:04
Calculating these values for higher concentrations produces the values shown in
Table 6.7.
On plotting the data (Fig. 6.32), it is evident that the optimal concentration that
minimizes the SEC occurs close to Cw = 35% and would require a system pressure
of 12,000 m 0.428 kPa/m = 5.14 MPa, which is 68% of the maximum working
pressure of the pipe. It would therefore be acceptable for these mildly abrasive
slurries. The conveying velocity would be 3.6 m/s. From the discussions in Sect.
6.3.3, it is also evident that this flow would likely be transitional.
204 6 Non-Newtonian Slurries and Suspensions
References
H.L. Lawler, N.T. Cowper, P. Pertuit, J.D. Tennant, Application of stabilised slurry concepts of
pipeline transportation of large particle coal. Presented at the 3rd international technical
conference on slurry transportation, Las Vegas, NV, 1978
J.L. Lumley, Drag reduction in turbulent flow by polymer additives. J. Polym. Sci. Macromol. 7,
263–290 (1973)
J.L. Lumley, Two-Phase Flow and Non-Newtonian Flow, in Turbulence, ed. by P. Bradshaw,
(Springer, Berlin, 1978)
V. Matoušek, A. Chryss, L. Pullum, Modelling the vertical concentration distributions of solids
suspended in a turbulent visco-plastic fluid. J. Hydrol. Hydromech. 69, 255–262 (2021)
A.B. Metzner, Flow of Non-Newtonian Fluids, in Handbook of Fluid Dynamics – Sec. 7, ed. by
V.L. Streeter, (McGraw Hill, New York, 1961)
A.B. Metzner, J.C. Reed, Flow of non-Newtonian fluids—Correlation of the laminar, transition, and
turbulent-flow regions. AICHE J. 1, 434–440 (1955)
M. Mooney, Explicit formulas for slip and fluidity. J. Rheol. 2, 210 (1931)
D.M. Newitt, J.F. Richardson, M. Abbot, R.B. Turtle, Hydraulic conveying of solids in horizontal
pipes. Trans. Inst. Chem. Eng. 33, 93–113 (1955)
L. Pullum, L.J.W. Graham, A new high concentration pipeline test loop facility. Presented at the
14th international conference on slurry handling and pipeline transport, Maastricht,
Netherlands, 1999
L. Pullum, L.J.W. Graham, The use of magnetic resonance imaging (MRI) to probe complex hybrid
suspension flows. Presented at the 10th international conference on transport and sedimentation
of solid particles, Wroclaw, Poland, 2000
L. Pullum, D.J. McCarthy, Ultra High Concentration and Hybrid Hydraulic Transport Systems in
Freight Pipelines, in Freight Pipelines, ed. by G.F. Round, (Elsevier Science Publishers,
Amsterdam, 1993)
L. Pullum, D.J. McCarthy, N.J. Longworth, Operating experiences with a rotary ram slurry pump to
transport ultra-high concentration coarse suspensions. Presented at the 13th international con-
ference on slurry handling and pipeline transport, Johannesburg, RSA, 1996
L. Pullum, L. Graham, P. Slatter, A non-Newtonian two layer model and its application to high
density hydrotransport. Presented at the 16th international conference on the hydraulic transport
of solids, Santiago, Chile, 2004
L. Pullum, L.J.W. Graham, M. Rudman, B. Aldham, R. Hamilton, ‘The ups and downs of paste
transport’. Presented at the 9th international seminar on paste and thickened tailings, Limerick,
Ireland, 2006
L. Pullum, P. Slatter, L. Graham, A. Chryss, Are tube viscometer data valid for suspension flows?
Korea-Australia Rheol. J. 22, 225–230 (2010)
L. Pullum, A. Chryss, L. Graham, V. Matoušek, V. Pĕník, Modelling turbulent transport of solids in
non-Newtonian carrier fluids applicable to tailings disposal. Presented at the 17th international
conference on transport and sedimentation of solid particles, Delft, The Netherlands, 2015
L. Pullum, D.V. Boger, F. Sofra, Hydraulic mineral waste, transport and storage. Ann. Rev. Fluid.
Mech. 58, 157–185 (2018)
B. Rabinowitsch, Ueber die viscosität und elastizität von solen. Zeitschrift Phys Chemie A 145,
1 (1929)
M.R. Rojas, A.E. Saez, Two-layer model for horizontal pipe flow of Newtonian and non-Newtonian
settling dense slurries. Ind. Eng. Chem. Res. 51, 7095–7103 (2012)
M. Rudman, H.M. Blackburn, L.J.W. Graham, L. Pullum, Turbulent pipe flow of shear-thinning
fluids. J. Non-Newtonian Fluid Mech. 118, 33–48 (2004)
S.N. Shah, D. Lord, Critical velocity correlations for slurry transport with non-Newtonian fluids.
AICHE J. 37, 863–870 (1991)
J. Singh, M. Rudman, H.M. Blackburn, A. Chryss, Pullum, Turbulent flow of non-Newtonian fluids
in a partially blocked pipe. Presented at the 19th Australasian fluid mechanics conference,
Melbourne, Australia, 2014
References 207
J. Singh, M. Rudman, H.M. Blackburn, A. Chryss, L. Pullum, L.J.W. Graham, The importance of
rheology characterization in predicting turbulent pipe flow of generalized Newtonian
fluids. J. Non-Newtonian Fluid Mech. 232, 11–21 (2016)
P. Slatter, The role of rheology in the pipelining of mineral slurries. Miner. Process Extr. Metall.
Rev. 20, 281–300 (2000)
P.T. Slatter, E.J. Wasp, Transition velocity estimation for visco-plastic fluids. Presented at the 13th
international conference on transport and sedimentation of solid particles, Tiblisi, Georgia, 2006
R.B. Spelay, Solids transport in laminar, open channel flow of non-Newtonian slurries. Disserta-
tion, University of Saskatchewan, Saskatoon, 2007
R. Stainsby, R.A. Chilton, Prediction of Frictional Pressure Losses in Laminar and Turbulent
Non-Newtonian Pipe Flows, in Pumping Sludge and Slurry, (Mechanical Engineers Publica-
tions Ltd, London, 1998), pp. 13–30
A.M. Talmon, W.G.M. van Kesteren, D.R. Mastbergen, J.G.S. Pennekamp, B. Sheets, Calculation
methodology for segregation of solids in non-Newtonian carrier fluids. Presented at the 17th
international seminar on paste and thickened tailings, Vancouver, Canada, 2014
A.D. Thomas, Pipelining of coarse coal as a stabilized slurry: Another viewpoint. Presented at the
4th international technical conference on slurry transportation, Las Vegas, USA, 1979
A.D. Thomas, K.C. Wilson, New analysis of non-Newtonian turbulent flow-yield-power-law fluids.
Can. J. Chem. Eng. 65, 335–338 (1987)
A.D. Thomas, K.C. Wilson, Rough-wall and turbulent transition analyses for Bingham
plastics. J. South. Afr. Inst. Min. Metall. 107, 359–364 (2007)
Z. Wang, P. Larsen, Turbulent structure of water and clay suspensions with bed load. J. Hydraul.
Eng. ASCE 120, 577–600 (1994)
K.C. Wilson, The dense-phase option for coarse-coal pipelining. J. Pipelines 2, 95–101 (1982)
K.C. Wilson, Modelling the effects of non-Newtonian and time-dependent slurry behaviour.
Presented at the 10th international conference on the hydraulic transport of solids in pipes,
Innsbruck, Austria, 1986
K.C. Wilson, A.D. Thomas, A new analysis of the turbulent-flow of non-Newtonian fluids.
Can. J. Chem. Eng. 63, 539–546 (1985)
K.C. Wilson, A.D. Thomas, Analytic model of laminar-turbulent transition for Bingham plastics.
Can. J. Chem. Eng. 84, 520–526 (2006)
Chapter 7
Vertical and Inclined Slurry Flow
7.1 Introduction
The previous chapters have dealt with horizontal flows, this chapter considers other
configurations important in practice, namely, vertical and inclined flows. From a
practical point of view, the most important effects of flow inclination are those on the
pressure drop and the deposition limit velocity. If compared with the horizontal flow,
the modification of the pressure drop is due primarily to the developed additional
static pressure drop. The effects on the frictional pressured drop and the deposition
limit velocity are related to the flow stratification, which also varies with the flow
incline.
Vertical flow is of particular importance in the mining industry. The ultimate
application of hydraulic hoisting is the proposed mining of deep ocean polymetallic
nodules, which are up to 50 mm in size and are found at depths up to 5000 m. A
typical d/D ratio is of the order of 10-1 (van Wijk et al. 2016).
Inclined flow is also of significant interest. Pipeline systems that transport settling
slurries often contain inclined sections of various lengths and slopes. Relatively
short lengths of pipe on an upward incline commonly occur at the inlet end of
transport systems. Examples include in-plant transfer lines—for example, in the oil
sands mining industry (Sanders et al. 2004)—and dredging installations with
inclined suction pipes of trailing hopper dredges and cutter dredges.1 The dredge
suction pipes can be more than 100 m long and set to a broad range of angles (van
den Berg 2013). Lengths of pipe at both positive and negative inclinations may be
incorporated in long-distance product pipelines running over rugged terrain (Thomas
and Cowper 2017).
From a practical point of view, the most important effects of flow inclination are
those on the pressure drop and the internal structure of slurry flow (the flow
1
Note that where possible, pipelines in plants should be either sensibly horizontal or vertical to
minimize the extra losses associated with inclined flows.
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 209
R. Visintainer et al., Slurry Transport Using Centrifugal Pumps,
https://doi.org/10.1007/978-3-031-25440-6_7
210 7 Vertical and Inclined Slurry Flow
stratification). If compared with the horizontal flow, the modification of the pressure
drop is due primarily to the developed additional static pressure and its sensitivity to
slurry density. The modification of the internal structure depends primarily on
particle size and results in a variation of friction loss and deposition limit velocity.
weight of the added solids, i.e., by ρwg(Ss - Sf)CviΔz. The resident (in situ) solids
concentration (Cvi) is reflected in the potential energy gain. It is applied to the height
difference (Δz), and for a pipe at an upward inclination (θ), the height difference per
unit length of pipe equals sinθ. For the pipeline in Fig. 7.1a, the total pressure drop
between A and C is a sum of the frictional pressure drops in the horizontal,
ascending, and descending sections of the pipeline. We will see that the frictional
pressure gradient can differ between these sections if settling slurry is transported.
In an inclined pipe, as in Fig. 7.1b, the total pressure drop is composed of the
frictional part and the static part, and it is expressed using the Bernoulli equation as
follows:
Δp
= im,fric,θ ρw g þ Smi ρw g sin θ ð7:1Þ
L
where im,fric,θ is the frictional dimensionless pressure gradient in the inclined pipe,
L is the length of the inclined pipe = (Δz/sinθ), and Smi is the relative in situ density
of the mixture. Smi = ρmi/ρw and is given by
Here Ss represents the relative density of the solids and Sf the relative density of
the carrier fluid. The pressure needed to lift the particles depends on the in situ
concentration (Cvi), but predictive pressure drop formulae are usually based on the
delivered concentration (Cvd). Determining the difference between Cvi and Cvd due
to phase slip (hold-up) is therefore an important step in pressure drop evaluation for
inclined flows.
In a vertical flow, the slip can be extreme. It is easy to imagine a vertical flow
heavily loaded with particles (requiring a large pressure differential simply to hold
them in place), but operating at a small enough velocity that very little in the way of
solids throughput is delivered. This particle hold-up effect can be important for large,
heavy particles (i.e., those that give fully stratified flow in horizontal pipes), but it is
usually less significant in heterogeneous flow and is negligible in pseudo-
homogeneous flow (where the in situ concentration is virtually the same as that
delivered).
Vertical flow is simply a special case for sinθ = 1.0. Consider a height (Δz) of a
vertical riser, with pressure p1 at the bottom and p2 at the top, giving a net pressure
force in the upward direction of ( p1 - p2)πD2/4. The weight of slurry is a downward
force which can be written (ρmigΔz)πD2/4 or (ρwgSmiΔz)πD2/4. The remaining force,
directed downward as it resists motion, is based on the shear stress at the pipe wall
(τo), multiplied by the wall area (πDΔz). The resulting force balance is given by
πD2 πD2
ðp1- p2 Þ = ρw g Smi Δz þ τ0 πDΔz ð7:3Þ
4 4
or
212 7 Vertical and Inclined Slurry Flow
p1 - p2 4τ
= ρw g Smi þ 0 ð7:4Þ
Δz D
For Eq. (7.4), the wall shear stress must be determined according to friction
conditions in the vertical flow of slurry, which will differ from the horizontal models
discussed earlier. Options for these vertical models are discussed in Sect. 7.4. The
relative density (Smi) is based on Cvi, while the delivered concentration (Cvd) is a
more common input parameter to slurry flow calculations. In vertical flow with
particles uniformly distributed throughout the pipe cross section, the difference
between Cvi and Cvd can be expressed using the ratio of the flow velocity and the
particle settling velocity. The average upward velocity of the liquid is denoted Vf,
and the particles are falling relative to Vf, with hindered settling velocity (vt′), so that
their net upward velocity is Vf – vt′. If Cvd is thedelivered
volumetric solids fraction,
then from continuity, the solids flow rate Cvi V f- v0t πD2 =4 equals CvdVmπD2/4.
Usually the difference between Vf and the average upward velocity of the mixture
(Vm) can be neglected, and the continuity can be written as follows:
Cvd
C vi = ð7:5Þ
1 - vt 0 =V m
The ratio vt′/Vm (slip velocity/mean velocity), which appears in Eq. (7.5), can be
very important in vertical slurry transportation systems. For a typical sand-water
slurry with an average particle diameter of 0.6 mm, the equivalent value of vt is about
0.05 m/s. As the hindered values of settling velocity (vt′) will be less than vt, we can
see that in many cases of practical interest (for which Vm will be 1.5 m/s or greater),
the ratio vt′/Vm will be significantly less than 0.1. This ratio can then be neglected in
the term (1 – vt′/Vm), so that to a reasonable approximation, Cvi = Cvd. Note that
transient conditions with a low loading of particles (i.e., vt′ ≈ vt) may be a deter-
mining factor for the design velocity.
Because fluid velocity goes to zero at the wall, teetering may occur where solids
are held up in the region near the wall of the vertical riser, consequently influencing
the size distribution of the delivered solids. In coarser particle systems, particles may
migrate away from the wall and form a high-concentration region in the middle of
the pipeline section, with water rushing along the side; see Fig. 7.2.
Vm
7.3 Internal Structure of Flow 213
1.0 1.0
sand 3 sand 3
0.8 ascending 0.8 descending
Vm = 3.5 (m/s) Vm = 3.5 (m/s)
0.6 0.6
y/D (-)
y/D (-)
0.4 0.4
0 deg 0 deg
0.2 25 deg 0.2 -25 deg
35 deg -35 deg
0.0 0.0
0.0 0.2 0.4 0.6 0.0 0.2 0.4 0.6
c (-) c (-)
1.0 1.0
sand 1 sand 1
0.8 ascending descending
0.8
Vm = 3.5 (m/s) Vm = 3.5 (m/s)
0.6
y/D (-)
0.6
y/D (-)
0.4 0.4
0 deg 0 deg
0.2 25 deg 0.2 -25 deg
35 deg -35 deg
0.0 0.0
0.0 0.2 0.4 0.6 0.0 0.2 0.4 0.6
c (-) c (-)
Fig. 7.3 Chord-averaged concentration distributions for two slurries (Sand 1 (fine) and Sand
3 (coarse)) at flow velocity Vm = 3.5 m/s in a 150 mm pipe at different pipe inclinations.
y/D = 0 is at the bottom of the pipe. (Matoušek 1996)
214 7 Vertical and Inclined Slurry Flow
1 1
Incline = ± 15⁰ up up
Incline = ± 25⁰
down down
0.8 0.8
y/D (-)
0.6
y/D (-)
0.6
0.4 0.4
broad-graded sand broad-graded sand
0.2 0.2 Vm = 2.5 m/s, Cvd = 0.24,
Vm = 2.5 m/s, Cvd = 0.24,
d50 = 0.55mm, D = 100mm d50 = 0.55mm, D = 100mm
0 0
0 0.2 0.4 0.6 0 0.2 0.4 0.6
c (-) c (-)
Fig. 7.4 Comparison of the solids distribution of broadly graded sand (d50 = 0.55 mm) in settling
slurry flow at Vm ≈ 2.5 m/s and Cvd ≈ 0.24 in a 100 mm pipe inclined to ±15° (left) and ± 25°
(right). (Matoušek et al. 2022)
1 1
Incline = +25⁰ Incline = -25⁰
0.8 0.8
layered model layered model
0.6 0.6 predictions
y/D (-)
y/D (-)
predictions
0.4 0.4
narrow-graded sand narrow-graded sand
0.2 Vm = 1.84 m/s, Cvd = 0.17, 0.2 Vm = 1.84 m/s, Cvd = 0.17,
d50 = 0.87mm, D = 100mm d50 = 0.87mm, D = 100mm
0 0
0 0.2 0.4 0.6 0 0.2 0.4 0.6
c (-) c (-)
Fig. 7.5 Comparison of the distribution of narrowly graded sand in a 100 mm pipe inclined at ±25°
with a layered model’s predictions. (Matoušek et al. 2019)
The observed difference in profiles for the stratified coarse flow results from two
factors. First is the different relative velocities between the sliding bed and the flow
above the bed in the ascending flow (slow sliding of the bed) and in the descending
flow (fast sliding of the bed). Second is the corresponding different development of
the shear layer at the top of the sliding bed. The same effect of flow inclination on
solids distribution was observed in an inclinable U-tube composed of 100 mm pipes
for settling slurry flow of broadly graded medium sand (Fig. 7.4). As in Fig. 7.3,
plots in Fig. 7.4 compare concentration profiles measured simultaneously in the
ascending and descending limbs of the U-tube. The flows of the same Vm and Cvd
differ significantly in the degree of stratification; the descending flow is fully
stratified, while the ascending flow exhibits a gradual change in local concentration
across the pipe’s entire cross section.
Mechanisms responsible for the observed effect are included in a layered model
for inclined settling slurry flows (Matoušek et al. 2018). Figure 7.5 shows the results
of the layered model, plotted together with measured concentration profiles for a
slurry of 0.87 mm sand in a 100-mm inclined U-tube. Both the measured and
predicted solids distributions exhibit a strong difference in stratification of positively
and negatively inclined flows set to 25°.
7.4 Vertical Flow Applications 215
p1 - p 2 V2
= Smi ρw g þ ρf m ð7:6Þ
Δz 2D
The question immediately arises as to what values to use for ρ and f in the final
term of Eq. (7.6). If settling slurries are transported vertically, then the equivalent
liquid approximation should be appropriate. Hence,
p 1 - p2 0 V 2m
= ρw g Smi þ ðSf þ A Cvd ðSs- Sf ÞÞf f ð7:7Þ
Δz 2gD
If the solid particles are small (less than approximately 150 μm), the coefficient
A’, associated with the equivalent liquid model, in Eq. (7.7) tends to 1.0 (i.e., the
mixture density is used in the last term in Eq. (7.6)).
Larger particles tend to migrate away from the pipe wall. When this occurs, the
near-wall zone, in which most of the velocity is gained, can be considered to behave
like a carrier fluid, and A’ tends to zero (i.e., the carrier fluid density is used in the last
term of Eq. (7.6)). Behavior differing from that just described can occur at very high
solids concentrations, which inhibit coring and force particles against the pipe wall
(see Wilson et al. 1979; Wilson 2004).
Coarse particles (larger than approximately 0.7 mm) exert collisional stresses
against a pipe wall which can contribute significantly to the total wall shear stress,
and hence to the frictional pressure drop, particularly at high Cvd (Matoušek 2009;
Shook and Bartosik 1994; Spelay et al. 2017). The friction factor based on one of the
models for the collisional stress should be used ( f = fcoll) together with ρ = ρf in
Eq. (7.6).
In a vertical slurry transportation system, Govier and Aziz (1972) proposed that
the minimum operating mixture velocity should exceed twice the terminal settling
velocity of the largest particles if there is a range of particle sizes. The settling
velocity in still water of industrially crushed mineral particles is normally much less
216 7 Vertical and Inclined Slurry Flow
(a) (b)
down
Vs ,up Vs ,down
up Cvi ,up Cvi ,down
Cvd < Cvi Cvd < Cvi
Cvd const.
Vm Cvi ,up Cvi ,down
2
Fig. 7.6 Flow conditions in vertical inverted U-tube: (a) average concentration and velocities
(schematic); (b) photo showing larger Cvi and smaller average solids velocity (Vs) in the left
upgoing limb than in the right downcoming limb of the U-tube
Dz
Dz
equal, allowing the shear stress term to be eliminated by combining the equations for
upward and downward flow.
In the U-tube, the pressure drops are measured by connecting the tappings to a
pair of differential pressure transducers with the connecting lines filled with liquid
(Fig. 7.7). Hence, the recorded pressure differences will be the manometric pressure
differences:
218 7 Vertical and Inclined Slurry Flow
and
Using these measured values, and substituting into Eq. (7.4), the result gives the
following expression for Cvi , the mean in situ solids concentration (i.e., the average
for riser and downcomer):
Δp12,man þ Δp43,man
C vi = ð7:10Þ
2ðSs - Sf Þρw gΔz
and
Δp12,man þ Δp43,man
Smi = þ Sf ð7:11Þ
2ρw gΔz
Although Cvi is usually very close to the delivered volumetric solids concentra-
tion (Cvd), the two are not identical. Clift and Clift (1981) derived an expression for
the fractional difference using the hindered settling velocity correlation of Richard-
son and Zaki (1954). The magnitude of the fractional difference between Cvd and Cvi
is generally less than (vt′/Vm)2, where vt′ is evaluated at the estimated delivered
concentration. It follows that the difference is negligible in almost all cases of
practical interest, implying that Eq. (7.11) can generally be used to calculate Cvd.
Krupička (2014) evaluated the accuracy of the inverted U-tube in determining
Cvd by comparing its results with other methods (e.g., sampling). The absolute value
of accuracy (standard deviation) of the concentration was approximately half a
percent for various solids distributions with particle sizes from 0.18 mm to 1.6 mm.
To design and operate successful and safe transport systems containing inclined
pipes, it is essential to predict accurately the effect of pipe inclination on pressure
drop and deposition limit velocity. This effect can be predicted by using semi-
empirical correlations or by computing a layered model.
7.5 Inclined Flow Applications 219
Compared with the horizontal case, flow up an incline tends to require higher
throughput velocities to avoid deposition. This tendency is of the greatest signifi-
cance for coarse-particle flows, which tend to stratify.
Durand’s predictive correlation for the deposition limit velocity in horizontal flow
introduced the dimensionless deposition limitffi velocity, now called the Durand
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
parameter, F L,m = V sm = 2 g ðSs =Sf - 1Þ D (see also Chap. 3). Wilson and Tse
(1984) employed the Durand parameter FL,m to express the effect of pipe inclination
on the deposition limit velocity. The effect was expressed as the change in FL,m and
presented in a chart relating ΔD with θ, where ΔD = FL,m,θ - FL,m, with FL,m,θ
evaluated for Vsm,θ in a pipe inclined to the angle θ from the horizontal, and FL,m for
Vsm in a horizontal pipe. Wilson and Tse based the chart on the results of their
experiments with coarse particles of sand and gravel (1.1, 1.7, 5.0, and 5.7 mm) in a
76-mm pipe that could be inclined up to 40° from the horizontal and on additional
literature data from a 55.8 mm pipe with the inclination angle varying from -20° to
+30°. They found that the velocity at the limit of deposition initially increases with
the angle of upward inclination, reaching a maximum when this angle is about 30°.
For the materials tested, this maximum velocity was approximately 50% larger than
that required to move a deposit in a horizontal pipe. This large difference is clearly
important for both design and operation of pipelines with inclined sections.
Figure 7.8 shows the experimentally observed ΔD plotted against θ. The designer
or operator concerned with coarse-particle transport in inclined pipes should take the
conservative approach of using the envelope curve for the experimental points,
shown as a dashed line in the figure. An appropriate approximation of the curve is
0:50ð0:6366θÞ2 0:02ð2:29θÞ2
given by ΔD = 0:75θ - for θ > 0, and ΔD = 0:75θ -
1 - 0:6366θ 1 - 2:29θ
otherwise, where θ is expressed in radians.
In applying this method, the deposition limit velocity is first estimated for the
horizontal case (Vsm), and then ΔD is obtained from Fig. 7.8 or the equations above
for the required angle of inclination θ. The deposition limit velocity in inclined flow,
Vsm,θ, is then calculated as:
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
V sm,θ = V sm þ ΔD 2gðSs =Sf - 1ÞD ð7:12Þ
This quantity gives the upper limit of stationary deposition when the pipe is
inclined at angle θ. An example of correcting Vsm for inclination angle will be shown
in the case study presented in Sect. 7.6.
The usefulness of the Wilson–Tse chart was verified experimentally for very
coarse slurries (gravel fractions with a majority of experimental data for d/D > 0.02)
by Spelay et al. (2016) and de Hoog et al. (2017). For finer slurries, however, the
chart was found to overestimate the effect of inclination on the deposition limit
velocity. Consequently, the Wilson–Tse chart can be considered as an upper limit,
giving the maximum inclination effect for different solids fractions in a pipe. This
limit is reached if solids are very coarse, and thus, the flow is virtually fully stratified.
For partially or fully stratified flows, it is appropriate to analyze the deposition
condition in an inclined pipe using the force balance on the bed (the lower layer of
the stratified flow). In essence, a layered model states that a deposit of solids must
move when the forces in the direction of flow exceed those resisting motion. The
principal effect of pipe inclination is to introduce an axial component of the
submerged weight of the bed. In ascending flow, this component increases the
force-resisting motion, accounting for the increase in the deposition limit velocity.
Interfacial conditions, namely shearing of the top of the bed, must also be taken into
account while formulating the bed force balance.
Figure 7.9 shows a generalized chart with curves for various d/D (particle size/
pipe size) produced by the layered model (Matoušek et al. 2019) including curves for
settling slurry flows that are not fully stratified. The calculations are done for one
pipe size (D = 100 mm) and one concentration (Cvd = 0.2). The model predictions
are compared with the original chart curve (calculated using the VSCALC program
given in Appendix A).
The model predictions show that values of ΔD are very small for fine particles of
d/D = 0.003, which agree with the experimental results by Spelay et al. (2016) for
the same d/D. Moreover, the predictions show that ΔD is quite sensitive to d/D for
medium particle sizes (in the range of 0.003 < d/D < 0.03), but considerably less
sensitive to coarser particles. For ascending flows (θ > 0°), the model curves reach
the original chart at d/D ≈ 0.04. For d/D > 0.04, model curves are slightly above the
original chart, but the sensitivity of ΔD, with respect to d/D, is low. Curves of the
generalized chart vary with D, but the variation is minor, and the trend holds. In
summary, the generalized chart gives a useful indication of the effect of the flow
7.5 Inclined Flow Applications 221
0.5
D = 100mm, rs = 2650 (kg/m3) 0.3mm dia
0.4
0.3 0.9mm dia
0.2 2.0mm dia
0.1
Δ D (-)
4.0mm dia
0.0
-0.1 6.0 mm dia
Fig. 7.9 Generalized Wilson–Tse chart predicted by a layered model for slurry flows of different
particle sizes. (Matoušek et al. 2019)
inclination on Vsm for different d/D smaller than 0.04, but for best results, Vsm in
inclined, partially stratified flows should be determined by a layered model compu-
tation for the particular flow condition rather than by using the chart.
In inclined slurry pipes, predicting the total pressure drop requires an evaluation of
both the static and friction components. The static component requires one to
distinguish between Cvd and Cvi when determining the weight of the slurry column
produced by the elevation change. The frictional component may vary with pipe
incline, and the effect of pipe inclination on friction loss depends on the type of
slurry flow. Friction of homogeneous flow is insensitive to pipe inclination. The
effect of pipe inclination on pseudo-homogeneous flow can be often neglected.
Heterogeneous and stratified flows, however, produce friction losses that are quite
sensitive to flow inclination.
Earlier chapters have described how the frictional solids effect (i.e., the frictional
gradient in excess of that for fluid alone) varies with the contact load, which trans-
mits its submerged weight to the lower part of the pipe via granular contact. When
the pipe is horizontal, all of the submerged weight of the contact-load solids is
transferred in this fashion, but when the pipe is inclined at some angle (θ) to the
horizontal, only the cross-pipe gravitational component g cos θ is involved. Thus, in
the absence of other changes, the solids effect component of the frictional gradient
222 7 Vertical and Inclined Slurry Flow
Fig. 7.10 Inclined pipe (schematic). Note that p2 is always taken downstream of p1, whether the
flow is inclined or declined. Gravity vector and the cross-pipe component of gravity vector are
indicated in red
will be multiplied by cos θ, whereas the fluid component will be unaffected by pipe
inclination. This principle is explored by both the semi-empirical and mechanistic
approaches to modeling of the inclination effect.
A widely used semi-empirical approach by Worster and Denny (1955) quantifies
the effect of an angle of pipe inclination (θ) on the total pressure drop (Δp = p2 – p1),
over the pipe length (L ) (see Fig. 7.10). It does this by using the total solids effect
(i.e., a difference in total pressure gradients for slurry and carrying liquid at the same
velocity in the same pipe). The technique is performed as follows. The solids effect
in inclined flow is expressed as a sum of the solids effects in corresponding
horizontal and vertical pipes, as shown in Fig. 7.10.
In a horizontal pipe, the solids generated dimensionless pressure gradient is
simply the frictional solids effect (im - if), which may be written Δi(0). In a horizontal
projection of an inclined pipe (as in Fig. 7.10), the solids effect is Δi(0)cosθ, and it is
considered as the frictional contribution by the cross-pipe component of the sub-
merged weight of particles to the total gradient by solids friction. Further contribu-
tion is made by the solids effect on the static gradient in the vertical projection of the
inclined pipe, and it is assumed to be (Ss - Sf)Cvd sin θ by Worster and Denny.
Hence, the phase slip is neglected, as is the frictional contribution of the solids in the
vertical section.
From the expression for the solids’ static contribution, it follows that the pressure
gradient in the Worster–Denny formula is actually related to the manometric pres-
sure gradient:
Δp12,man
= if þ Δið0Þ cos θ þ ðSs - Sf ÞC vd sin θ ð7:13Þ
ρw gL
Equation (7.13) can be used directly to express the gauge pressures in a sub-
merged pipe, such as the suction pipe of a dredge (see Case Study 7.2 in Sect. 7.6). It
gives the gauge pressure gradient between points 1 and 2, where the line gauge
pressures are expressed relative to the pressures outside the submerged pipe at points
1 and 2. The absolute static pressure gradient (i.e., the actual pressure gradient within
the pipe, including the static contribution from both the solids and the carrier fluid)
can be expressed as follows:
7.5 Inclined Flow Applications 223
p1 - p2
= if þ Δið0Þ cos θ þ Smd sin θ ð7:14Þ
ρw gL
where Smd is the relative density of the slurry based on the delivered concentration
(Cvd) and if is the frictional gradient due to the carrier fluid alone, which in inclined
flows is equal to if for horizontal flow. Note that while the values of p1 and p2 will
change depending on whether or not the pipe is submerged, the difference between
them will not change for a given inclination angle; hence, Eq. (7.14) holds true for
both submerged and un-submerged pipes.
The frictional pressure gradient is given by:
Δp12,fric
= if þ Δið0Þ cos θ ð7:15Þ
ρw gL
Hence, the Worster–Denny formula assumes that the frictional pressure drop is
the same in the ascending pipe and in the descending pipe of the same inclination
angle (cos(-θ) = cos θ).
Due to its assumptions, the Worster–Denny approach can only be successfully
applied in settling slurry flows where stratification is weak or nonexistent, and where
the effect of phase slip can be neglected.
For ascending stratified flows, the Worster–Denny equation may give low esti-
mates of the static part of the total pressure drop since it uses Cvd instead of Cvi,
which ought to be used in the static pressure-drop term. It may also give conserva-
tively large estimates of the inclination effect on the frictional pressure drop in an
ascending pipe, as it does not account for the reduced stratification that can occur as a
result of bed shear in inclined flows (Figs. 7.3, 7.4 and 7.5). In a few cases, such as in
Fig. 7.11b below, the two contradictory trends affecting the total pressure drop can
compensate each other so that a successful prediction of the manometric pressure
gradient is reached for ascending partially stratified flow (but not for descending
flow).
0.5 0.5
Manometric gradient (a) (b)
Frictional gradient
0.4 0.4
Cvd
Worster-Denny mano, grad. by Cvd
Worster-Denny mano. grad. by Cvi
Cvi
0.3 Worster-Denny frictional gradient 0.3
0.2 0.2
im (-)
im (-)
0.1 0.1
0 0
broad-graded sand narrow-graded sand
-0.1 Vm = 2.5 m/s, Cvd = 0.24, -0.1 Vm = 1.84 m/s, Cvd = 0.17,
d50 = 0.55mm, D = 100mm d50 = 0.87mm, D = 100mm
-0.2 -0.2
-50 -40 -30 -20 -10 0 10 20 30 40 50 -50 -40 -30 -20 -10 0 10 20 30 40 50
q (deg) q (deg)
Fig. 7.11 Dimensionless pressure gradient at various angles of inclination for (a) broadly graded
sand and (b) narrowly graded sand. Comparison is made with predictions by Worster and Denny.
(Matoušek et al. 2022)
224 7 Vertical and Inclined Slurry Flow
The recent experiments with settling slurries of various medium and medium-to-
coarse sands in an inclinable, inverted U-tube at the Institute of Hydrodynamics in
Prague showed considerable differences in the frictional head loss between the
ascending limb and the descending limb of the U-tube when inclined from zero to
±45° from the horizontal (Fig. 7.11). Due to the significantly stronger stratification
(as in Fig. 7.4, which shows results of the same experiment as Fig. 7.11a, and in
Fig. 7.5, for the same experiment as in Fig. 7.11b), the descending flow exhibits
more resistance than the ascending flow at the same average velocity and delivered
concentration if the flow is not very steep. This is particularly the case at mild slopes
around ±15°, where it may lead to a substantially larger frictional gradient in the
descending flow than in the ascending flow. Note that gamma ray measurements of
Cvi in the ascending and descending pipes have been used to calculate the frictional
gradient from the manometric gradient measurements in Fig. 7.11.
Based on this experimental data, it is suggested to restrict the use of models that
apply the simple correction of cos(θ) to the solids’ frictional head loss (such as in the
Worster–Denny formula) to ascending flows, if the flow is stratified (including
partially stratified flows). This finding—that descending flows at rather mild pipeline
slopes exhibit greater resistance than do the corresponding ascending flows—has
important implications for slurry systems that contain upward-inclined portions and
transport solids that tend to deposit on the pipe invert. These systems normally need
to be arranged for local emptying through descending gravity flow after intentional
or unintentional stops. The increased resistance of the descending flow, as compared
with the pumped ascending flow, must be taken into account in predicting the
gravity flow.
A layered model, which considers the existence of a shear layer at the top of the
bed, can recognize variations in the stratification and slip-based differences between
Cvi and Cvd due to the inclination of the settling slurry flow. The model uses the
recognized conditions for predicting all types of the pressure drop (total, frictional,
manometric) in the inclined flow. Its present version (Matoušek et al. 2018) predicts
both ascending (Figs. 7.12 and 7.13) and descending (Fig. 7.13) flows. These flows
0.2
0.1 Experiment
Worster-Denny prediction
Layered model prediction
0
0 10 20 30 40 50
q (deg)
7.5 Inclined Flow Applications 225
0.40
narrow-graded sand
0.35 Cvd = 0.25, d50 = 0.87mm, D = 100mm
0⁰
0.30 -15 ⁰
0.25 15 ⁰
25 ⁰
im.frict (-)
0.20
35 ⁰
0.15 45 ⁰
0.10
0.05
0.00
0 1 2 3 4 5 6
Vm (m/s)
Fig. 7.13 Dimensionless pressure gradient predicted by the layered model for a narrowly graded
sand at different inclinations. (Matoušek et al. 2018)
include those which are at least partially stratified, where the entire bed slides in the
direction of the flow (the limitation for ascending flows) but are restricted to flows
where the sliding bed is slower than the flow above the bed (the limitation for
descending flows). Three-layer models (e.g., Doron and Barnea 1993) are required
to account for flows where one layer exceeds the flow above the bed.
As shown in Fig. 7.12, the Worster–Denny prediction of the frictional gradient is
larger than that seen in the experimental data for the medium-to-coarse sand. The
relatively low friction is due to increased shearing in the bed layer as a result of the
slip between the bed and flow above it, resulting in reduced stratification in the
ascending pipe (and a corresponding decrease of the normal force of the bed on the
pipe wall), compared to the corresponding horizontal pipe case. The layered model
can account for this reduction in stratification (see Fig. 7.5), and as Fig. 7.12 shows,
it gives more accurate predictions in this case.
The unified layered model also solves inclined flow with a stationary deposit in
the pipe. Note that contrary to the sliding bed, the stationary deposit does not
contribute to the static pressure drop, as the deposit transfers its weight completely
to the pipe wall (Jovanović and Matoušek 2019).
As a concluding remark, it should be emphasized that in many cases, the friction
loss is only a small part of the total pressure drop over an inclined pipe. The static
component dominates if the inclination is sufficiently steep. In the cases shown in
Fig. 7.11, the static component starts to dominate at angles larger than approximately
10° and smaller than approximately -25°.
When using the 4CM to predict friction losses for inclined flows, it is
recommended to only apply an inclination angle correction to the horizontal friction
loss predictions for the heterogeneous and stratified components, and to assume that
the friction losses for the pseudo-homogeneous and carrier components will be
226 7 Vertical and Inclined Slurry Flow
independent of the inclination angle. If the flows are known to be strongly stratified,
it is recommended to use a layered model instead of the 4CM to predict the pressure
drop, especially in mildly descending flows. This approach will provide a more
representative model of the physics at play and will accordingly predict the pressure
drop for these types of flows more accurately.
In an iron ore mine, the ore (ρs = 4950 kg/m3 as in Table 2.4) is ground to <100 μm
in a subsurface facility and then pumped vertically 800 m to the surface through a
steel 8-inch pipe (D = 202.7 mm, commercial steel, ε = 45 μm). The carrying liquid
is water at 10 °C, and the delivered concentration of transported solids Cvd = 0.20 at
the slurry velocity Vm = 2.0 m/s. Determine the pressure requirement to pump the
slurry to the surface. Assume also an emergency condition at which ore particles
cannot be processed in the subsurface facility, and solids of 15 mm in size must be
lifted in the pipeline at the same slurry velocity and pressure requirement as for the
normal condition.
(a) Pressure Requirement to Transport Processed Ore
As determined by Eq. (7.7), the pressure requirement is dependent on both the
static conditions (primarily the specific gravity of slurry in the vertical pipe (Smi))
and on the friction conditions (namely A’ for the equivalent liquid model). If the ore
is ground to <100 μm (dmax = 0.1 mm), then the slip of solids relative to fluid is
negligible, as confirmed by Eq. (7.5), Cvi = 1 -Cvtvd0 =V m = 1 - 0:005=2
0:2
= 0:2005, in which
the terminal settling velocity of a non-spherical ore particle is determined by
Eq. (2.67) as vt = 0.013 m/s for ψ = 0.79, and is converted to the hindered settling
velocity (vt′) using Eqs. (2.71) and (2.72), assuming Cv = Cvd. Given the close
correspondence of Cvi and Cvd, we can proceed with the assumption Cvi = Cvd, and
therefore Smi = Smd = Sf + (Ss - Sf) Cvd = 1.00 + (4.95 - 1.00) 0.20 = 1.79.
The friction loss is predicted by the equivalent liquid model (Eq. (7.7)), with a
constant A’ = 1 for the fine ore. Hence:
0 V 2m
Δp = Δzρw g Smi þ ðSf þ A C vd ðSs - Sf ÞÞf f =
2gD
h i
4
= 800 9810 1:79 þ ð1 þ 1 0:2 3:95Þ0:016 = 14:27 MPa
19:62 0:2027
0:156
Cvi = = 0:2033, Smi = 1:803, and:
1 - 0:465=2
h i
4
Δp = 800 9810 1:803 þ 0:016 = 14:28 MPa
19:62 0:2027
Note that since the delivered concentration (which is what the pump will see) has
been reduced significantly (from 0.2 to 0.156), the power absorbed by the pump will
also be reduced, although not as much as the concentration ratio would indicate, due
to a likely increase in the pump head reduction factor (see Chap. 9). If pump speed
can be increased to allow operation at higher velocity (higher flow rate), then
applying the higher velocity criterion Vm,req = 2vts would actually be beneficial to
the vertical transport of this coarse ore. Using Vm = Vm,req = 2 ‧ 1.286 = 2.57 m/s
leads to an increase of the delivered concentration (Cvd) to 0.164 due to lesser slip.
The new condition is Cvi = 0.2006, Smi = 1.792, vt′ = 0.469 m/s, and
Δp = 14.27 MPa. If additional pump speed increase is possible, the velocity can
be increased until the full pump power is absorbed, further minimizing slip and
maximizing delivered concentration.
Case Study 5.3 dealt with the transport of coarse gravel in a horizontal discharge
pipeline (D = 489.0 mm) connected with a cutter suction dredge (CSD). The CSD
dredges the coarse gravel from the seabed and conveys it at a slurry flow rate (Qm) of
1.13 m3/s and delivered concentration (Cvd) of 0.215 through the pipeline system,
228 7 Vertical and Inclined Slurry Flow
which comprises both the discharge line and the suction line. In this study, the
suction line (located on the dredge ladder) is considered. It carries slurry from the
cutter head to the centrifugal pump on board the dredge. Assume that the centerline
of the pump suction is at the same height as the seawater level. The suction pipe has a
diameter (D) of 539.8 mm (commercial steel, ε = 45 μm) and a length of 9 m. The
inclination angle of the suction pipe can be adjusted according to the depth to be
dredged, but the current depth is 5.8 m.
Procedure As discussed in Case Study 5.3, the flow of coarse gravel is stratified.
Hence, it would be appropriate to use a layered model to predict both the deposition
limit velocity and the pressure drop in the inclined suction pipe. For convenience,
simpler semi-empirical methods will be used to estimate the effect of flow incline on
both the deposition limit velocity and the pressure drop.
(c) Deposition Limit Velocity for Coarse Gravel in the Suction Pipe
The inclination angle (θ) of the suction pipe is given by the flow depth
zdepth = 5.8 m and the pipe length L = 9 m as sinθ = zdepth/L = 5.8/9 = 0.644,
and thus θ = 0.700 radians ≈ 40° to the horizontal. The dimensionless ΔD for the
effect of θ on the Vsm is estimated from the Wilson–Tse chart (Fig. 7.8) or from the
equation included in VSCALC (see Appendix A):
The deposition limit velocity (Vsm) would be 3.41 m/s if the suction pipe were
horizontal (Eq. 4.14) for μs = 0.48). When the pipe is inclined to +40°, the
deposition limit velocity (Vsm,θ) is given by Eq. (7.12):
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
V sm,θ = V sm þ ΔD 2gðSs =Sf - 1ÞD = 3:41
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
þ0:346 19:62ð2:65=1:026 - 1Þ0:5398 = 4:83 m=s
Compared with the horizontal pipe, the deposition limit velocity is larger by more
than 40% in the pipe inclined to +40°, and it is very close to the flow velocity
Vm = 4Qm/(πD2) = 4.94 m/s. However, the ΔD estimation may be too conservative
using Fig. 7.8 or the corresponding equation for d50/D = 10.7/539.8 = 0.020 (see
Fig. 7.9, where the blue curve corresponding to d50/D = 0.02 falls below the Wilson
and Tse curve). If estimated from Fig. 7.9, then ΔD ≈ 0.27. Thus, Vsm,θ ≈ 4.5 m/s,
which implies that the suction pipe is free of deposit during the operation. In Case
Study 5.3, a possible lowering of Vm to 4.8 m/s was discussed for the discharge line,
primarily for wear reasons. This lowering would result in the velocity of 3.94 m/s in
the suction line, which is far below the Vsm,θ of 4.5 m/s, so it can no longer be
considered as a viable option.
7.6 Case Studies 229
Δp12,man
= if þ Δið0Þ cos θ þ ðSs - Sf Þ C vd sin θ
ρw gL
= 0:030 þ ð0:204 - 0:030Þ cos 40 þ ð2:65 - 1:026Þ 0:215 sin 40
= 0:388 ðm of standard water=m of pipeÞ
On multiplying this by the pipe length L = 9 m, the manometric head over the
submerged pipe is found to be 0.3889 = 3.49 m of water. The greater part of this
head arises from lifting the particles through the suction pipe from the bottom to the
water level. This is represented by the static term containing sinθ on the right-hand
side of the equation: (2.65–1.026)0.215sin409 = 2.02 m of water.
Note that the spatial volumetric concentration (Cvi) should be used instead of Cvd
to determine the static head and that the Worster–Denny assumption Cvd = Cvi might
not be appropriate for the coarse transport under consideration. Perhaps the slip is
not negligible, resulting in Cvi > Cvd and causing the static part of the head (hence,
the total pressure drop) to be underestimated by the calculations above. The detailed
calculation would require the layered model.
(e) Static Pressure at the Suction Side of the Dredge Pump
To determine the static pressure at the pump suction side, the Bernoulli equation
is applied to the suction pipe inlet (1) and outlet (2), which is also the suction side of
the pump. The conditions at (1) are as follows: the flow velocity is zero, and the
gauge pressure ( p1) equals the hydrostatic pressure exerted by the seawater column
of the height (zdepth), where ρp1g = Sf zdepth . The gauge pressure head (in meters
w
standard water) at (2) is then given by the Bernoulli equation:
p2 V2
= - ð Smi - Sf Þzdepth þ Smi m þ H frict þ H minor
ρw g 2g
P V 2m
Substituting zdepth = L sin θ, and including fitting losses H minor = Smd K 2g
with an estimated total loss coefficient ∑K = 1.5, gives:
X V2
p2
= - ð Smd - Sf ÞL sin θ þ Smd 1 þ K m
þ if L þ Δið0ÞL cos θ
ρw g 2g
Substituting Smd and the previously calculated manometric head from Worster–
Denny (which includes both static and friction lost contributions) gives:
" #
p2 Δp12,man X V2 4:942
=- þ Smd ð1 þ KÞ = - ½3:49 þ 1:375ð1 þ 1:5Þ
m
ρw g ρw g 2g 19:62
= - 7:77m of standard water:
The pressure at the suction inlet of the pump (which is located at the same height
as the water line) is subatmospheric to the extent of about 8 m of water. The absolute
p p þ p2
static pressure head is 2,abs = atm ≈ 10 - 7:77 = 2:23 m of water. It depends
ρw g ρw g
on the properties of the dredge pump and whether such a pressure is sufficiently high
to avoid cavitation in the pump (Cavitation is discussed in Chaps. 8 and 10).
By using the delivered concentration (Cvd and Smd) in our calculations, we are
neglecting the slip between solids and liquid in the suction line. If significant, the
pump inlet pressure ( p2,abs) will be lower than calculated, due to the higher in situ
concentration and resulting higher static head caused by slip. If the pressure ( p2,abs)
is insufficient to prevent cavitation, then either the slurry flow rate (Qm) or the
delivered concentration (Cvd) must be lowered to increase the pressure at the pump
inlet. If Qm is lowered from 1.13 to 1.08 m3/s, resulting in a flow velocity of 4.72 m/s
in the suction line (closer to Vsm,θ ≈ 4.5 m/s), then the pressure head at the pump
suction inlet (relative to atmospheric) increases to -7.38 m of water and the absolute
pressure head ( p2,abs) to 2.62 m, which provides a greater margin against cavitation.
(f) Dredging of Finer Gravel Using the CSD System
Let us consider that finer (predominantly heterogeneous) material is dredged by
the same CSD system. The material is the one used in Case Study 5.2(b): density (ρs)
of 2650 kg/m3; d50 = 3.5 mm; d85 = 7.2 mm; d50h = 3.76 mm; dmax = 13 mm;
Xf = 0.05; Xp = 0.10; Xh = 0.75; Xs = 0.10 (d50h, Xh, and Xs modified for
D = 539.8 mm). For convenience, we consider the same dredging conditions as
for the coarser gravel operation: zdepth = 9 m; angle of incline θ = +40°;
Qm = 1.13 m3/s; Cvd = 0.215.
References 231
In the discharge line, the operating velocity of 6.0 m/s resulting from
Qm = 1.13 m3/s is substantially above the Vsm of 4.07 m/s (by the VSCALC) and
sufficiently above Vmin of 4.85 m/s (by the 4CM). In the suction pipe, the horizontal
Vsm = 4.29 m/s (by the VSCALC) and ΔD ≈ 0.10 for d50/D = 3.5/539.8 = 0.006
from Fig. 7.9. Thus:
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
V sm,θ ≈ 4:29 þ 0:10 19:62ð2:65=1:026 - 1Þ0:5398 ≈ 4:7 m=s
This is also sufficiently below the operating velocity of 4.94 m/s; therefore, the
suction pipe is deposit-free.
Due to lower friction losses (im,0 = 0.110 at Vm = 4.94 by the 4CM) than for the
coarser gravel, the manometric head over the suction pipe is smaller:
Δp12,man
= 0:030 þ ð0:110 - 0:030Þ cos 40 þ ð2:65 - 1:026Þ 0:215 sin 40 = 0:316
ρw gL
To summarize, the finer gravel (d50 = 3.5 mm) can be dredged at the same
conditions as the previously considered coarser gravel (d50 = 10.7 mm) by the cutter
suction dredge system. The finer gravel operation ensures better margins against
deposition in the suction pipe and against cavitation in the dredge pump.
References
R. Clift, D.H.M. Clift, Continuous measurement of the density of flowing slurries. Int. J. Multiphase
Flow 7, 555–561 (1981)
E. de Hoog, M. in ‘t Veld, J.M. van Wijk, A.M. Talmon, An experimental study into flow assurance
of coarse inclined slurries. Presented at the 17th International Conference on Transport and
Sedimentation of Solid Particles, Prague (2017)
P. Doron, D. Barnea, A three-layer model for solid-liquid flow in horizontal pipes. Int. J. Multiphase
Flow 19, 1029–1043 (1993)
J. Evans, C.A. Shook, Dispersion and slip effects in hydraulic hoisting of solids. Can. J. Chem. Eng.
69, 1166–1173 (1991)
G.W. Govier, K. Aziz, The flow of complex mixtures in pipes (Van Nostrand Reinhold, 1972),
p. 792
T.W. Hagler, Means for Determining Specific Gravity of Fluids and Slurries in Motion (U.S, Pat,
1956), p. 2678529
M. Jovanović, V. Matoušek, The effect of longitudinal slope on solids transport and friction in
particle-laden flow above stationary deposit in pipe. Presented at the 19th International Confer-
ence on Transport and Sedimentation of Solid Particles, Cape Town (2019)
232 7 Vertical and Inclined Slurry Flow
J. Krupička, Mathematical and physical modelling of pipe flow of settling slurries (Dissertation,
Czech Technical University in Prague, Prague, 2014)
V. Matoušek, Internal structure of slurry flow in inclined pipe. Experiments and mechanistic
modelling (Presented at the 13th Intl. Conf. on Slurry Handling and Pipeline Transport,
Johannesburg, RSA, 1996)
V. Matoušek, Pipe-wall friction in vertical sand-slurry flows. Particulate Sci. Tech. 27, 456–468
(2009)
V. Matoušek, J. Krupička, M. Kesely, A layered model for inclined pipe flow of settling slurry.
Powder Tech. 333, 317–326 (2018)
V. Matoušek, M. Kesely, Z. Chára, Effect of pipe inclination on internal structure of settling slurry
flow at and close to deposition limit. Powder Tech. 343, 533–541 (2019)
V. Matoušek, Z. Chára, J. Konfršt, Settling slurry transport – Effects of solids grading and pipe
inclination, in Slurry Technology – New Advances, ed. by T. Jones, (IntechOpen, London, 2023)
J.F. Richardson, W.M. Zaki, Sedimentation and fluidisation: Part I. Trans Instn. Chem. Engrs. 32,
35–53 (1954)
R.S. Sanders, J. Schaan, R. Hughes, C.A. Shook, Performance of sand slurry pipelines in the oil
sands industry. Can. J. Chem. Eng. 82(4), 850–857 (2004)
C.A. Shook, A. Bartosik, Particle-wall stresses in vertical slurry flows. Powder Tech. 81, 117–124
(1994)
R.B. Spelay, R.G. Gillies, S.A. Hashemi, R.S. Sanders, Effect of pipe inclination on the deposition
velocity of settling slurries. Can. J. Chem. Eng. 94, 1032–1039 (2016)
R. Spelay, R.G. Gillies, S.A. Hashemi, Kinematic friction of concentrated suspensions of neutrally
buoyant coarse particles (Presented at the 20th Intl. Conf. on Hydrotransport, Melbourne, 2017)
A.D. Thomas, N.T. Cowper, The design of slurry pipelines – Historical aspects (Presented at the
20th Intl. Conf. on Hydrotransport, Melbourne, 2017)
C.H. van den Berg, IHC Merwede Handbook for Centrifugal Pumps and Slurry Transportation
(MTI Holland, Kinderdijk, 2013)
J.M. van Wijk, A.M. Talmon, C. van Rhee, Stability of vertical hydraulic transport processes for
deep ocean mining: An experimental study. Ocean Eng. 125, 203–213 (2016)
K.C. Wilson, Energy consumption for highly-concentrated particulate slurries (Presented at the
12th Intl. Conf. on Transport and Sedimentation of Solid Particles, Prague, 2004)
K.C. Wilson, J.K.P. Tse, Deposition limit for coarse particle trqansport in inclides pipes (Presented
at the 9th Intl. Conf. on the Hydraulic Transport of Solids in Pipes, Rome, 1984)
K.C. Wilson, N.P. Brown, M. Streat, Hydraulic hoisting at high concentration: A new study of
friction mechanisms (Presented at the 6th Intl. Conf. on the Hydraulic Transport of Solids in
Pipes, Canterbury, 1979)
R.C. Worster, D.F. Denny, Hydraulic transport of solid materials in pipelines. Proc. Instn. Mech.
Engrs. 169, 563–586 (1955)
Chapter 8
Centrifugal Slurry Pumps
8.1 Introduction
The first half of this book concentrated on the behavior of various kinds of slurries in
pipes. Beginning with the current chapter, the second half of the book deals with
other components of a pipeline transport system, and with how those components
interact. We begin with an overview of the hydraulic performance and mechanical
design of centrifugal slurry pumps, with emphasis on the information required by the
pump user rather than the pump designer. The hydraulic analysis is based on the
classical case of the flow of a simple liquid. Topics covered include the calculation of
pump head, power and efficiency, and the concept of Net Positive Suction Head
(NPSH), which is introduced in association with the phenomenon of pump cavita-
tion. Various differences between slurry flow and liquid flow are mentioned, but it is
not until Chap. 9 that a complete treatment of the effect of solids on pump head and
efficiency is presented. The scaling of pump performance is covered in detail,
including several case studies. The importance of pumps of this type as prime
movers for high-tonnage slurry operations was shown in Chap. 1; other types of
prime movers, such as positive displacement pumps, are referred to where appro-
priate, but their operation is not analyzed in detail in this book. After examining the
effects of slurries on pump performance in Chap. 9, Chap. 10 analyzes the ever-
important interactions between pump and pipeline systems, and Chap. 11 expands
on this topic to include other components of the slurry pipeline system and various
operational cases of practical interest. Pump and pipeline testing are discussed
further in Chap. 12, and Chap. 13 addresses the basics of erosive wear. Finally,
Chap. 14 considers the total cost of ownership and best practices for slurry pump
selection.
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 233
R. Visintainer et al., Slurry Transport Using Centrifugal Pumps,
https://doi.org/10.1007/978-3-031-25440-6_8
234 8 Centrifugal Slurry Pumps
We begin the analysis of pumps by considering the flow of a simple liquid of density
ρf, which moves through the pump at a discharge flow rate denoted Q, measured in
volume per unit time. The pump adds energy to the fluid, enabling it to move through
the piping, overcoming both friction and the effects of any elevation changes.
Chapter 2 showed that this energy represents the sum of the pressure energy,
potential energy, and kinetic energy; and Bernoulli’s equation expresses each of
these three energy terms as heights of the flowing liquid, (see Sect. 2.4.2). This
concept of height, or head, is also used in defining the energy added by the pump, as
defined by Eq. 2.17 in terms of pressures, velocities, and elevations at pump entry
(A) and exit (B). This equation is repeated below in Eq. (8.1). According to common
convention, the term HPUMP from Chap. 2 is replaced here, and throughout the
remainder of the text, by a simple H. In the past, pump head was sometimes called
the “total dynamic head,” or TDH, but this term is no longer widely used.
V 2B - V 2A pB - pA
H = þ þ ðzB - zA Þ ð8:1Þ
2g ρf g
The power output of the pump is determined by the mass flow rate and the head,
and is given by:
This relation applies in any consistent system of units. Thus, for SI units, Eq. (8.2)
gives the power output in Watts, which is usually divided by 1000 to obtain
kW. (In US customary units, a common form of this equation includes a numerical
coefficient to account for the inconsistent units: Pout = Q H Sf/3960, where Pout is
expressed in horsepower, Q in gallons per minute, and H in feet, and Sf is the specific
gravity of the fluid being pumped.)
The input power to the pump is Tω, where T is the torque on the pump shaft and ω
is the angular velocity. With seconds as the unit of time, ω is equal to 2πn, where n is
measured in revolutions per second. In practice, the speed of rotation is usually
measured in revolutions per minute, even where SI units are generally employed.
The symbol N will be reserved for revolutions per minute (rpm). Thus:
2πN
ω = 2πn = ð8:3Þ
60
2πNT
Pin = 2πnT = ð8:4Þ
60
8.1 Introduction 235
With torque in Newton-meters, this equation gives power in Watts. Finally, the
ratio of the pump’s output-to-input power is its efficiency, denoted η, (not to be
confused with the usage of this symbol for slurry viscosity):
Pout
η= ð8:5Þ
Pin
For an ideal pump, the efficiency (η) is 1.00 or 100%. Although real pumps
necessarily have lower values, efficiencies of over 90% can be achieved for large
pumps. Efficiencies tend to be somewhat lower in slurry applications, for reasons to
be discussed in Chap. 9. In the past, the term “efficiency” was sometimes used to
represent the ratio of actual to ideal values of quantities other than power. This usage
can lead to confusion and will be avoided here. In this book, efficiency refers only to
the power ratio defined by Eq. (8.5).
Efficiencies can be obtained from pump tests, which will be described in
Chap. 12. If a given pump is driven at a constant shaft speed (i.e., a fixed N ), a
series of readings of Q, H, and T can be obtained at various openings of a throttling
valve located downstream of the pump. The head is plotted directly against discharge
flow rate, as shown in Fig. 8.1. This curve is known as the head-discharge charac-
teristic, or may also be called the head-flow, head-quantity, or head-capacity curve,
or simply the H-Q curve. The required power (Eq. 8.4), and the efficiency (Eq. 8.5),
are also plotted against Q, as shown in the figure. The limiting suction performance
curves (to be discussed in Sect. 8.3) can also be plotted.
100
Head BEP
H (m)
80
HBEP
60
Efficiency
h (%)
P (kW)
40 1200
Power 800
20
QBEP 400
0 0
0.0 0.5 1.0 1.5 2.0
Q (m3/s)
With N constant, the efficiency (η) varies only with the ratio HQ/T, where T is
always greater than zero. Thus, η will be zero at the no-flow condition (Q = 0) and
again when the H-Q curve intercepts the flow rate axis at H = 0. Between these
extremes, the efficiency curve displays a maximum, as shown in the figure. This
maximum defines the best efficiency point, or BEP, and the associated flow rate and
head are often identified as QBEP and HBEP.
The curves shown in Fig. 8.1 refer to a single angular velocity, or pump rotary speed
(N ), and if the tests were repeated at a different speed, all the points would shift. This
behavior can be plotted as a series of H-Q curves for various speeds, together with
contour lines of constant efficiency, power, and NPSHR (defined later; see Sect. 8.3)
to produce a multispeed pump performance chart, as shown in Fig. 8.2. Test data are,
in general, not required for each curve; instead, the various constant speed curves are
constructed based on simple scaling relations. All flow rates (including both QBEP
and the flow rate at H = 0) shift in direct proportion to N, while all heads (including
both the no-flow head and HBEP) shift in proportion to N2.
Similar shifts are found if the shaft speed is kept constant but the pump’s size
scale is changed while maintaining geometric similarity (i.e., where all dimensions
scale in the same ratio). The dimension customarily used to describe this ratio is the
outer diameter of the impeller vanes, denoted for this purpose by D (not to be
confused with the pipeline diameter). It is known that when the size of the pump
as a whole is altered, heads scale in proportion to D2, and flow rates scale in
proportion to D3. The theory underlying the scaling relations will be presented in
Sect. 8.2.
It follows from the scaling relations that the curves for all members of a set of
geometrically similar1 pumps operating at various N values can be collapsed onto a
single curve by using an appropriate system of dimensionless axes. One system of
this type (White 2016) uses as abscissa the dimensionless discharge flow rate Q/nD3,
with n in revolutions per second. (The quantity is dimensionless only in consistent
units; e.g., Q in m3/s and D in meters, or Q in ft3/s and D in feet). The efficiency (η) is
already dimensionless, and the dimensionless head takes the form gH/n2D2 (again in
consistent units).
Using this dimensionless axis system, the head and efficiency characteristics of a
representative pump are plotted in Fig. 8.3. Despite differences in pump size and
angular velocity, all the points for geometrically similar pumps will lie on a single
line of gH/n2D2 versus Q/nD3. When the pump efficiency is plotted against Q/nD3,
the same basic uniformity occurs; however, in this case, the points do not always lie
on a single line. This occurs because smaller pumps often have a reduced efficiency
due to the effects of surface roughness in the hydraulic passages, which often does
not scale according to geometric similarity but remains relatively constant across a
range of pump sizes. Furthermore, the ratio of mechanical friction power (from
bearings and seals) to total pump power usually increases as pump size decreases.
Empirical and semi-empirical formulas are available for estimating these effects.
(See, for example, Anderson 1980; ANSI/HI 14.6 2022; White 2016; Yoda et al.
2016). A reduction in efficiency with pump speed may also occur, as the pump
power reduces at a greater rate than that of the mechanical friction power. This effect
is again most noticeable in smaller pumps. It is often measured during pump
prototype testing and incorporated into the multispeed pump performance chart
(Fig. 8.2).
The dimensionless characteristic curves shown in Fig. 8.3 display representative
centrifugal pump behavior. The peak of the efficiency curve is gently rounded and
operation at flow rates slightly above or below the peak does not carry a large
1
In this context, “geometrically similar” means that all dimensions of one pump in the set can be
related to all dimensions of another by a fixed ratio. Thus, the set of geometrically similar pumps all
represent “pure” scales of each other.
238 8 Centrifugal Slurry Pumps
7.0 1.40
gH
6.0 n2D2 1.20
n2D2
gH
5.0 1.00
BEP
4.0 0.80
3.0 h 0.60
h
2.0 0.40
1.0 0.20
0.0 0.00
0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14
Q
nD3
penalty. The dimensionless H-Q curve is rather flat near the no-flow condition (i.e.,
at low Q/nD3), but it droops increasingly downward as the BEP point is approached.
The well-known affinity laws for pumps represent the scaling relations mentioned
above and are closely associated with dimensionless plots such as Fig. 8.3. The
affinity laws are used for point-to-point conversion of test data for one pump size and
shaft speed in order to plot characteristic curves for some other size or speed. These
laws apply only to the members of a set of geometrically similar machines. Since
data points for any member of the set fall onto the same curve when plotted in the
dimensionless axis system, it follows that they can also be plotted directly onto the
curve for another member of the set. Thus, each data point from a test of a pump with
impeller diameter D1, run at angular speed n1, can be transferred to the curve of a
different pump with diameter D2 and speed n2. Although the affinity laws can be
expressed in various ways, the system used here is consistent with that of the
McGraw-Hill Pump Handbook (Cooper 2001) and also with the affinity laws for
slurry pipelines mentioned in earlier chapters.
The affinity law for discharge flow rate ensures that the scaled point has the same
value of Q/nD3 as the test point, i.e.,
3
n2 D2
Q2 = Q1 ð8:6Þ
n1 D1
Simultaneously, the scaled head H2 is obtained from the test point value H1 by
means of:
8.2 Hydraulic Design and Specific Speed 239
2 2
n2 D2
H2 = H1 ð8:7Þ
n1 D1
This affinity law for head ensures that the value of gH/n2D2 is maintained. The
gravitational acceleration has been canceled from the equation. Also, the problem of
units is sidestepped, as the affinity laws apply in any system of units. In practice, N1
and N2 in revolutions per minute are usually used in place of n1 and n2 in revolutions
per second.
The required power depends on the product of H and Q, and it scales according to
the relation:
3 5
n2 D2
P2 = P1 ð8:8Þ
n1 D1
This scaling relationship for power is not as accurate as those for flow and head,
since the pump efficiency may vary slightly with pump size or speed, as mentioned
previously. The inaccuracies are generally small (less than a few percent), especially
for larger slurry pumps (D > 1 m), but for accurate performance prediction,
designers usually make corrections based on testing or experience.
It is particularly important to note that discharge flow rate, head, and power must
be scaled together. A change of size affects flow rate, head, and power, but all in
different proportions, in accord with the three affinity laws. To obtain, for instance,
the H-Q curve for a scaled-up pump, each of the test data points must have both
H and Q scaled by the affinity laws. The new characteristic curve is then drawn
through the series of scaled-up points.
The scaling concept has been used in several sections in previous chapters for the
scale-up of friction loss from one pipe size to another on a point-by-point basis. The
simplest example was for the laminar flow of non-Newtonian slurries, for which the
affinity laws were given by Eqs. 6.15 and 6.16. In a more general sense, the use of
scaling relations forms a recurring theme in this book.
The previous section treated the centrifugal pump as a system component or “black
box” without examining the shape of the flow passages or the fluid mechanics of the
pumping process. This section addresses these physical considerations in greater
detail.
240 8 Centrifugal Slurry Pumps
A centrifugal pump has two main components. The first is the rotating element,
which comprises the shaft and the impeller, which includes the vanes that act on the
fluid. The second component is the stationary element, which consists of the casing
that encloses the impeller, together with the associated stuffing boxes and bearings.
The hydraulic design of a centrifugal pump is concerned with dimensioning the
impeller and the casing to provide the required performance characteristics. Any
pump design usually has more than one combination of component dimensions that
can be arranged to give a specified performance. The combination selected will
depend on the intended application and on any hydraulic or mechanical limitations.
Slurry pumps face several additional challenges, including the need to pass large
solids, the requirement for a robust rotating assembly (because the slurry density
exceeds that of water), and the need for thicker sections to withstand the effects
of wear.
Centrifugal pumps for the general industry exist in a wide variety of arrangements
to suit different applications. Pumps may contain multiple stages of impeller and
casing; various suction configurations such as in-line, double suction, or self-
priming; impeller profiles from radial to mixed flow to axial propeller; volute-spiral
casings with diffuser vanes; and additional passages or structures for the balancing of
hydraulic loads.
Slurry pumps, on the other hand, show less variety. They are normally single-
stage, end-suction type, with radial or mixed-flow impellers, using simple vaneless
casings of the concentric or semi-volute type, and having a minimum of specialized
hydraulic balancing structures. Representative sections are shown in Fig. 8.4.
Flow through pumps can be quite complicated. To aid in the description of these
flow fields, we must specify certain directions or coordinates. The axial direction is
parallel to the pump shaft and is positive in the direction of the fluid inflow (Fig. 8.4).
The radial direction is directly outward from the centerline of the shaft. The
tangential direction is perpendicular to both the axial and radial directions,
representing the tangent to the circular path of a rotating point. Points on the impeller
have only tangential velocity, given by ωr or 2πnr, where ω is angular velocity in
radians per second, n is in revolutions per second, and r is the radius from the shaft
centerline. A further direction, useful for defining the flow relative to the curvature of
the impeller passage, is the meridional direction. This direction lies within a plane
passing through the shaft centerline and follows the projection of the fluid stream-
lines onto this plane. Thus, the meridional direction has both radial and axial
components.
The meridional and tangential velocities are used to plot the velocity triangles at
the entry and exit of the impeller. The absolute velocity of the fluid (i.e., its velocity
measured relative to the ground) is denoted by c, with subscripts m and t denoting
components in the meridional and tangential directions. The absolute velocity of a
point on the rotating impeller, denoted by u, is necessarily in the tangential direction,
so a directional subscript is not required. Further, the subscripts 1 and 2 are used to
distinguish conditions at the entry and exit of the impeller, respectively.
The entry and exit velocity triangles are shown in Fig. 8.5. Note that the velocity
(w), which closes the vector triangle, represents the velocity of the fluid relative to
the impeller. As this velocity will (ideally) follow the inclination of the blades, the
angle (β) shown in the vector triangles will represent the blade angle (i.e., the angle
between the impeller blade surface and the tangential motion of the impeller at that
point). The exit blade angle (β2) is an important design parameter affecting the
developed head, and the entry blade angle (β1) is set to minimize energy loss due to
turbulence as the fluid enters the impeller.
The vector triangles provide the information required for solving the moment-of-
momentum equation. In its simplest form, applicable when conditions do not vary
with time, this equation states that the applied torque (T ) must equal the moment of
the net momentum flux passing through a stationary control volume. As the control
Fig. 8.5 Velocity triangles at impeller entry (a) and exit (b)
242 8 Centrifugal Slurry Pumps
volume is stationary (i.e., based on the ground, not the impeller), the velocities used
in calculating the momentum flux must also be ground-based, or absolute velocities.
(For this reason, absolute velocities were used for the vector triangles.)
Referring to variables displayed in Fig. 8.5, the following relation can be derived
for the torque (T) required to maintain the rotation of the impeller:
Multiplying Eq. (8.9) by ω gives the power, and on dividing both sides of the
resulting equation by Q and g, we obtain the following relation for the head produced
by the impeller:
where cm2 is the meridional component of outlet velocity (directed radially outward
for most slurry pumps), which, in turn, is given by the discharge flow rate (Q)
divided by the exit area of the impeller, i.e.,
Q
cm2 = ð8:12Þ
π D2 b2
where b2 is the width between the side plates (or shrouds) at the outlet of the
impeller.
Equation (8.12), together with the evaluation of u2 as πnD2, can then be
substituted into Eq. (8.11) and the result combined with Eq. (8.10), where the final
term of that equation is ignored for the ideal case. The result forms the basic head
relation for the ideal pump, written:
gH i Q D2
= π2 - cot β2 ð8:13Þ
n2 D22 n D32 b2
The ratio D2/b2 and the blade angle β2 will both be constant for all members of a
set of geometrically similar pumps. Thus, the head-capacity curve for any set of
8.2 Hydraulic Design and Specific Speed 243
Fig. 8.6 H-Q curve obtained by subtraction of hydraulic losses from the ideal line
idealized pumps will give a single straight line when plotted on the dimensionless
axis system of Fig. 8.3. The analysis given here is central to the theoretical demon-
stration of the validity of the affinity laws of Eq. (8.6) to Eq. (8.8). Real H-Q
characteristics lie below the theoretical straight line, approaching it only at flowrates
near QBEP. However, conditions near this point are of greatest practical interest. For
all pumps of a given geometry, the BEP point will lie at a single location on the
dimensionless axis system of Fig. 8.3, but this location shifts for different
geometries.
The volute or casing of a pump has the task of converting the kinetic energy of the
fluid leaving the impeller into pressure energy. In an idealized pump, no losses are
considered to occur in either the casing or the impeller. In practice, hydraulic losses
due to viscosity (friction losses) and local turbulence (shock losses) occur in all the
wetted passages of the pump. The head-capacity curve of an actual pump results
from the subtraction of these losses from the idealized pump characteristic, on the
basis that friction losses increase in a linear fashion with flowrate, and there is only a
single flow rate for which the shock losses at the impeller inlet approach zero. An
example is shown in Fig. 8.6.
Curves of this type can be expected to be geometrically similar for all members of
a set of pumps with a given specific speed, which is defined in the next section. Thus,
one machine can be used as a model for all members of the set, with the affinity laws
providing the basis for transferring the results from the model to other machines. If
model testing is carried out accurately on pumps that are geometrically similar,
performance can generally be predicted within 2%.
244 8 Centrifugal Slurry Pumps
As it happens, for most pump geometries of practical significance, the BEP points lie
within a relatively narrow band on the dimensionless plot. Although values of the
two dimensionless coordinates are required to define the BEP point precisely, a
single parameter that gives the position along the efficiency point band can do almost
as well. This parameter is the dimensionless specific speed (ns), obtained by manip-
ulating the ratios that form the axes of Fig. 8.3 to eliminate the impeller diameter.
The definition of ns used by White (2016) is:
pffiffiffiffi
n Q
ns = ð8:14Þ
ðgH Þ3=4
where n is in revolutions per second, Q in m3/s, and H in meters. Both Q and H are
defined at the BEP point. A similar dimensionless variable, based on angular speed
(ω) rather than n, is used by Cooper (2001). In engineering practice, gravitational
acceleration is often omitted, giving the customary specific speed, denoted Ns and
defined by:
pffiffiffiffi
N Q
Ns = ð8:15Þ
H 3=4
When model tests are not available, empirical methods must be used instead. In this
case, pump design is usually carried out in two stages. The first stage uses dimen-
sionless performance coefficients for the main dimensions of the hydraulic compo-
nents, such as the impeller and the casing. The hydraulic contours of the components
are obtained in the second stage by using certain methods such as velocity triangles,
area progressions, slip estimates, etc. This allows the head and capacity performance
to be checked against requirements. In modern design practice, numerical simula-
tions of flow (CFD) are commonly used to refine designs, but the value of the
empirical methods in developing an initial geometry remains.
Stepanoff (1957) provided empirical relations and coefficients for calculating
impeller diameter and outlet width, casing throat areas, and other key dimensions for
water pumps based on known (or assumed) values of the specific speed, which may
be itself calculated from the desired best efficiency point conditions of flow rate,
8.2 Hydraulic Design and Specific Speed 245
0.6
Head coef. y
0.5
0.4
0.3
20 30 40 50 60 70 80
Ns
Fig. 8.7 Head coefficient at QBEP as a function of specific speed for a selection of modern slurry
pump designs
head and rotational speed, (Eq. 8.15). The specific speed may then be correlated to
important dimensionless variables, for example, the head coefficient (ψ = gH/u22),
which is itself proportional to the dimensionless ratio gH/n2D2, (Fig. 8.3). Based on
the correlated head coefficient, the desired impeller diameter may then be estimated.
The head coefficients at QBEP for a selection of modern slurry pump designs
covering a range of sizes, specific speeds, and wear configurations from light to
heavy-duty, are shown in Fig. 8.7. The correlations against specific speed and
number of vanes cover the practical range of application for most slurry pumps.
Note that as the number of vanes and the outlet blade angle (β2) increase, the amount
of head produced increases as well. Vane number for slurry pump impellers com-
monly ranges from three to five, although some impellers have as few as one, and
others have as many as seven. Values of β2 usually range from 15° to 35°.
Slurry pumps require thick sections for wear resistance and larger flow passages
that are capable of clearing large solids. As a result, their coefficient values differ
from those of water pumps. It is likely, for example, that a slurry pump will require
larger impeller outlet widths than a water pump, which will tend to shift the pump
QBEP to a larger value. However, this shift can be compensated by changes in the
geometry of the casing, as shown by Worster (1963), since the impeller and casing
work together to determine both QBEP and HBEP. For this reason, slurry pump
impeller and casing geometries are often mismatched, as far as their own QBEP is
concerned, relative to conventional water pump design practice, and the empirical
methods widely used in designing clear fluid pumps are not always applicable to
246 8 Centrifugal Slurry Pumps
slurry pumps, or else they must be modified according to experience. Casing and
impeller wear must also be considered. Depending on the operating circumstances
and the best economic compromise between efficiency and wear, the hydraulic
contour of the casing may follow various shapes: volute, semi-volute, or concentric.
This point is discussed in greater detail in Chap. 13.
In the idealized case, an impeller with a large number of infinitely thin, friction-
less vanes would produce the highest efficiency in a pump. In practice, for a water
pump, this number of vanes would range from five to nine. In a slurry or sewage
pump, it may be reduced to three or four in order to pass coarse solids and
accommodate extra vane thickness. Fewer vanes result in a steeper H-Q curve and
some reduction in efficiency. The number of vanes to be used must be decided after
considering the size of solids to be passed, the vane thickness, the location of the
inlet, and the overall design of the vane shape. To hold the efficiency loss to a
minimum, the vanes should have the correct inlet angle for shock-free entry of the
fluid at the design flow rate (which is usually the desired QBEP). In addition, the
outlet angle must be set to give the desired performance, and the shape between the
inlet and outlet must be selected to minimize the rate of change of velocity.
In practice, the meridional section of the impeller, combined with the location of
the inlet edge, almost always imposes a flow across this edge that is a combination of
axial and radial motion. As the tangential velocity at the inlet edge will therefore
vary, the inlet angle that gives shock-free entry will also vary along this edge. This
results in a vane geometry that is “twisted” from inlet to outlet, rather than the more
easily produced, and once common, “radial” vane geometry, where a fixed cross-
section of the vane is maintained from suction to hub shroud when viewed in the
radial plane. In large pumps, the effect on performance can be significant, but where
the pump size is small and the pump-specific speed is low, the disadvantage of
having a radial vane may be acceptable, and the higher manufacturing costs of
twisted vanes may outweigh their advantages. Slurry pump designs vary as a result
of the intended service, the manufacturer’s design expertise, and even the vintage of
the design. In very general terms, we can say that older slurry impeller designs have
simpler “square” meridional sections with radial vanes that are relatively easy to
produce.
The suction performance of a pump is also strongly influenced by the location of
the impeller vane inlet edge in the cross-meridional section and by the inlet area.
Placing the vane inlet at a small radius (closer to the shaft centerline) can give good
suction performance, but if the radius is too small, the vane thickness can “choke the
eye,” creating excessive velocities that lead to efficiency losses and increase the
chance of cavitation. Conversely, location of the vane inlet at a large radius increases
its tangential velocity, which can also increase the chance of cavitation by increasing
the local pressure drop. Thus, a balance must be maintained.
Meridional sections are shown schematically in Fig. 8.8, where the symbol RV
refers to radial vanes. More modern, low-to-medium specific speed designs (labeled
ME in the figure) often retain simple near-square meridional sections but employ
modern twisted vanes. Higher specific speed and/or lighter service impellers are
often of the type labeled HE.
8.2 Hydraulic Design and Specific Speed 247
Slurry pump casings vary from the true volute, water pump “T” type shown in
Fig. 8.8 to the semi-volute “C” type, to the annular (or concentric) “A” type.
Each combination of the types illustrated in Fig. 8.8 has its own hydraulic
performance and wear characteristics. For example, the HE/T combination generally
has the highest efficiency but is not necessarily the most forgiving for wear, whereas
the ME/C combination is capable of respectable efficiency while having a more
predictable wear performance. The wear of various types of casings and impellers is
discussed further in Chap. 13.
The maximum efficiency obtainable from properly designed water pumps is
given in the literature, for example by Anderson (1980), Karassik et al. (2008),
and ANSI/HI 20.3 (2020). Slurry pumps rarely achieve these values because of their
larger mechanical losses (due to relatively larger shafts and bearings) and the
hydraulic compromises necessary to improve wear life. With the use of the latest
design technology, including numerical methods of hydraulic design (see Chap. 13),
it is now possible to get to within a few percent of those values and still have a slurry
pump which gives good wear performance.
Figure 8.9 compares the optimum efficiency expected of water pumps according
to Anderson (1980) with the efficiency of actual slurry pumps covering a range of
sizes, specific speeds, and wear configurations from light to heavy-duty. The bulk of
designs fall within 2–8% of the optimum, with a mean difference of -4%.
As indicated in earlier chapters, pumps must be selected by matching their H-Q
performance to the requirements of the piping system. Fig. 2.19 of Example 2.A
shows that the intercept of the pump characteristic with the system characteristic
defines the operating head and flow rate. For settling slurries in particular, the
selection of appropriate operating conditions raises special considerations, to be
discussed in Chap. 10. Pump selection and cost considerations will be covered in
248 8 Centrifugal Slurry Pumps
15%
Percentage of pumps
10%
5%
0%
-15 -10 -5 0 5
h ACTUAL - h OPTIMUM
Fig. 8.9 Difference between optimum water pump and typical slurry pump efficiencies
Chap. 14. In the case of long pipelines, the total system head will be more than a
single pump can handle, and it becomes necessary to use several pumps in series.
This case is considered in Chap. 11, as is the less common case of several pumps in
parallel.
Both the hydraulic designer and the pump user may be interested in modifying a
pump’s behavior by “trimming” the impeller (i.e., reducing the impeller vane
diameter only) rather than using a scaled-down version of the complete pump that
maintains geometric similarity. The effect of trimming can be obtained from
Eq. (8.13). As the velocity u2 varies in direct proportion to the impeller diameter,
the ratio gH/n2D2 retains its significance, so that the scaling law for head (Eq. 8.7)
applies in the usual fashion. However, a different scaling law is required for
calculating the discharge flow rate. Assuming that the trimming operation involves
a relatively minor change in diameter, the impeller outlet angle (β2) and the impeller
outlet width (b2) remain relatively constant (Lazarkiewicz and Troskolanski 1965),
and the discharge flow rate ratio in Eq. (8.13) becomes Q/nD2 rather than Q/nD3
(as would be the case for geometrically similar impellers). On this basis, Q for a
trimmed impeller is proportional to D2, equivalent to reducing the exponent of the
diameter from 3 to 2 in Eq. (8.6). An alternative relation for flow rate that is in
customary use, particularly in the US, employs an empirical proportionality between
Q and D (Kittredge et al. 2008; Stepanoff 1957). In this case, the trimming ratio
behaves similarly to a speed ratio. Neither treatment is exact, with the former
overestimating and the latter underestimating the required diameter. In practice,
empirical adjustments are applied to both relations to reach an essentially equivalent
result. One such method is shown in Case Study 8.3 at the end of this chapter. For
8.3 Cavitation and Net Positive Suction Head 249
In most hydraulic applications, the gauge pressure (i.e., the pressure above the local
atmospheric pressure) is used. Negative gauge pressure (suction) is often found at
the pump inlet, and in some circumstances can produce the condition known as
cavitation. Cavitation occurs when the local pressure somewhere within the pump
(usually at the vane inlet) falls below the vapor pressure of the liquid (i.e., the
pressure at which it boils), designated pv. The vapor bubbles or “cavities” that form
are carried through the pump to an area of higher pressure where they collapse. The
cavities result in some blockage of the flow, reducing pump efficiency and devel-
oped head. Their collapse, if much vapor is present, can be relatively violent, leading
to vibration, noise, and, in severe cases, surface damage to the impeller or even the
casing. Eventually, the pump may cease pumping altogether. Severe surging and
water hammer may also result (Carstens and Hagler 1964). For all liquids, the vapor
pressure increases with temperature, so cavitation is more likely to occur, for the
same pump suction pressure, when the liquid is hot. When the vapor pressure is
expressed in terms of head of the liquid in question and is evaluated using the liquid
density at the operating temperature, it is known as the vapor pressure head (hv).
Figure 8.10 shows values of hv for water as a function of temperature, based on the
data from Table 2.1.
The local pressure at some point within the impeller may be lower than the
pressure at the suction inlet pipe. Therefore, to ensure cavitation-free operation,
every centrifugal pump requires that the absolute pressure at the suction inlet
exceeds pv by a certain margin. By convention, this margin is expressed in terms
of head of the liquid being pumped and is known as the net positive suction head or
NPSH, defined as the difference between the head produced by the pump and the
vapor pressure head at any point in the flow. When calculated at the pump inlet, it is
commonly referred to as NPSHA (“NPSH available”) and is given by:
V2
NPSHA = ðha þ H s Þ þ - hv ð8:16Þ
2g
where ha is the atmospheric head (that is, the head equivalent of atmospheric
pressure), and Hs is the pressure head at the pump inlet in units of gauge pressure
(that is, relative to atmospheric pressure). Both are expressed as head of the liquid
being pumped. The velocity head term V2/2g is also evaluated at the pump inlet.
If the suction pressure head (Hs) has been measured, Eq. (8.16) gives NPSHA
directly. Otherwise, Hs must be calculated by applying Bernoulli’s law to the suction
piping system, allowing for head losses due to friction and entry effects. The NPSHA
250 8 Centrifugal Slurry Pumps
1.2
1.0
0.8
hV (m)
0.6
0.4
0.2
0.0
10 20 30 40 50
Temperature oC
Fig. 8.10 Vapor pressure head of water as a function of temperature
often varies with the flow rate due to losses in the suction piping that may be
calculated directly from the piping geometry and the known or assumed friction
coefficients. Note that Hs is not the same as the difference in level from the liquid
surface in the sump to the eye of the pump, since the suction piping losses must be
considered.
If the NPSHA falls below the minimum NPSH required by the pump, cavitation
will occur. Obviously, the minimum NPSH required to avoid cavitation is an
important factor in selecting a pump or designing a piping system. Although a
well-designed pump usually has good suction performance, the minimum required
NPSH cannot be predicted with precision and must usually be measured for a pump
prototype. In determining the onset of cavitation experimentally, it is usual to carry
out tests where the suction pressure head (Hs) is reduced while the shaft speed and
flow rate are maintained constant. The water temperature is monitored, and the
appropriate value of hv is used in calculating NPSHA.
The test is repeated at different flow rates to determine the complete performance
characteristic of the pump. As the magnitude of Hs is decreased from an initial high
value, there is at first no change in the head produced by the pump. However, as
significant cavitation develops, the pump head diminishes. The NPSH value at which
this occurs is called the NPSHR for “NPSH required.” For most industrial pumps,
8.3 Cavitation and Net Positive Suction Head 251
22
84 320
h (%)
80
21
Efficiency
310
76
20
19
300
Head
18
H (m)
17 290
3% head drop
16
280
15
kW
14
Power 270
13
12 260
0 5 10 15 20
NPSHR 3% NPSHA (m)
s
0.10
0.08
0.06
0.04
0.02
0.00
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6
Q / QBEP
and vibration. Cavitation at this level is usually only detectable by direct observation
in the test lab, using a pump with clear windows and a strobe light to freeze the
movement of the vane. As such, many centrifugal pumps operate with some small
cavitation present. Fortunately, this poses little concern for typical industrial appli-
cations and even less for slurry applications, where erosive wear dominates and
pump mechanicals are built to withstand the higher forces of slurry transport.
The concept of specific speed can also be adapted to NPSHR performance by
defining a “suction specific speed” in which the head (H ) in Eq. (8.14) or Eq. (8.15)
is replaced by the NPSHR at QBEP. For flow rates away from QBEP, a pump’s suction
performance can be described using σ, a dimensionless cavitation parameter defined
as the ratio of NPSHR to the total head produced by the pump at the same flow rate.
NPSHR
σ= ð8:17Þ
H
While pump rotational speed influences NPSHR, for a pump of a given design,
the value of σ depends only on the ratio Q/QBEP. An example curve of this type is
plotted in Fig. 8.12.
Design factors affecting the NPSHR performance of typical slurry pumps include
the impeller inlet diameter and area; the shape, number, and inlet angle of the
impeller vanes; the area between vanes; the pump-specific speed; and the shape of
the suction passages. Figure 8.13 shows the cavitation parameter (σ) for a selection
of modern slurry pump designs covering a range of sizes, specific speeds, and wear
configurations from light to heavy duty.
Many older slurry pump designs will show higher NPSHR performance than
indicated in Fig. 8.13. This is also true for smaller, elastomer-lined impeller designs,
since these often use “square” meridional sections with a small inlet radius and
relatively thick radial vanes, both of which tend to affect suction performance
adversely. Slurry pumps, however, are generally designed to run at relatively low
speeds in order to reduce wear, so that in applications with sumps providing a
positive suction gauge pressure, the NPSHA is usually adequate for the pump,
8.4 Mechanical Design 253
1.00
0.10
0.01
10 100
Ns
Fig. 8.13 Cavitation parameter (σ) at QBEP as a function of specific speed for a selection of modern
slurry pump designs
regardless of the design. In suction pit and dredging services, this is less likely to be
the case; for these applications, the pump’s suction performance may have a great
effect on the system’s reliability and production rate. This point was shown in the
case study of a dredge ladder in Chap. 7, and it will be discussed further in Chap. 11.
It is recommended that every slurry pump application be checked to confirm that the
NPSHA is greater than the NPSHR. If tests for NPSHR have not been made with the
slurry of interest, it will be necessary to adapt the results of water tests by expressing
all heads, including the atmospheric and vapor pressure heads, as the height of slurry
rather than water. Further effects of slurries on NPSHR will be mentioned in
Chap. 11.
A centrifugal pump’s mechanical design must consider the loads generated by the
transmission of power from the driver, hydraulic reactions due to the inertial effects
of fluid motion and the pressure distributions within the pump passages, dynamic
response of the rotating assembly, and mechanical loads coming from the attached
piping. These must often be considered over a wide range of operating conditions
due to the nature of solids transport and mineral processing duties. We can divide the
loads on a pump assembly into two main categories. The first acts on the rotating
assembly, giving rise to shaft and bearing loads. The latter acts on the casing
assembly, either internally (via pressure) or externally from the connected piping.
254 8 Centrifugal Slurry Pumps
The loads can also be divided into hydraulic forces generated by pressure distribu-
tions within the pump, mechanical forces resulting from the transmission of force
between parts, or the mass and acceleration of the parts themselves.
The loads on the rotating assembly include:
• the torque load on the shaft,
• the hydraulic axial thrust due to the pressure imbalance between the suction and
drive sides of the impeller,
• the hydraulic radial load at the impeller, caused by variations in the pressure
distribution within the casing,
• mechanical radial loads due to imbalance of the rotating parts and coupling forces
at the drive end,
• the weight of the rotating parts, which may act radially or axially depending on
the pump’s orientation.
The torque load during operation can be calculated from the input power and shaft
speed using Eq. (8.4). However, the shaft must usually be designed for a torque
larger than the operating value, reflecting the fact that the motor is usually oversized
to some extent, so that higher torques can occur during transients such as start-up.
The hydraulic axial and radial loads can be sizable, often dwarfing the mechanical
loads of weight, imbalance, and coupling forces. These are calculated by the
designer using empirical methods based on test data and, more recently, numerical
modeling (CFD). They can change dramatically with operating conditions and may
increase at duties where the mechanical loads are decreasing. Also, since these loads
depend on pressure distributions, they can often be modified by hydraulic structures
within the pump, such as clearing vanes in the side gaps between the impeller and
the casing, or in the pressure balancing passages. However, these hydraulic struc-
tures, which are so useful for controlling loads (and therefore often used in clear fluid
pumps) must be applied with care in slurry pumps, as they may have negative
consequences with regard to wear performance.
The radial load due to impeller imbalance is estimated from the pump speed and
the balancing limits that are allowed on the impeller and shaft. In slurry pumps,
which are usually of low specific speed and overhung configuration, static balancing
is generally sufficient, provided that the impeller is not excessively wide in relation
to its diameter. The load due to misalignment of a coupling or V-belt drive depends
on the size and type of installation and is usually best estimated from recommenda-
tions of the supplier of the drive.
Impeller attachment by trapezoidal screw threads (such as the 29° Acme type) has
generally been found best capable of carrying the heavy loads required. The threaded
connection can be manufactured economically in the hard materials most commonly
employed in impellers and is protected from direct contact with the slurry flow,
unlike the keyed shaft and nut often used in clear liquid pumps. The large loads
associated with heavy-duty service require roller bearings, often with a separate
roller thrust bearing. Designs using double-row, taper-roller bearings with a preset
internal clearance are also common. Single-row, deep-groove ball bearings, often
8.4 Mechanical Design 255
used in process pumps, are only suited for light service slurry pumps. Two repre-
sentative bearing assemblies are shown in Fig. 8.14.
A conventional packed stuffing box is still the simplest and most common
rotating assembly seal. Configurations are similar to those used for water pumps,
with the lantern ring supplying a clear water flush to the center of the packing for
minimum dilution, or to the product side for maximum life. An expeller-type seal is
popular where dilution of the product is unacceptable or seal water is unavailable.
They are usually limited to single-stage pump applications, and they may suffer from
accelerated wear in slurries with a high concentration by volume of solids, or high
solids specific gravity. They also absorb power, which results in efficiency losses of
3% or more.
Mechanical seals are also used with slurry pumps, although much less often than
with process or water pumps. Designs targeted to slurry service were introduced in
the 1980s, and most applications focused on eliminating seal water and used single-
face designs with no quench or flush. Installations were generally limited to light-
duty applications or chemical process slurries such as potash and aluminum
processing, where diluting the slurry with seal water is costly. Seal designs have
improved over time, and double-face seals with a pressurized barrier fluid are
gaining acceptance in slurry service. One motivation has been the increasing cost
(including environmental considerations) of supplying seal water to pumps that
serve long-distance pipelines with concentrated mine tailings. Mechanical seal
designs for slurry service are constantly developing. A good reference for the latest
design and application information can be found in the Rotodynamic Centrifugal
Slurry Pumps standard ANSI/HI 12.1–12.6 (2021).
256 8 Centrifugal Slurry Pumps
We turn next to the materials and construction of the wearing or “wet end” parts.
For very abrasive slurries, hard metal is common, principally high chrome and
chrome molybdenum white iron alloys of 600–750 Brinell hardness. Single-wall
designs with the main casing in hard metal are simple to construct and maintain, and
are therefore common. These can be used unlined, or they can have liners of bonded
rubber or other material.
For slurries of an average particle size of less than 3 mm (1/8 inch) and a top size
of less than 12 mm (1/2 inch), natural rubber, synthetic elastomers, and urethanes,
collectively known as elastomers, are often used and can provide excellent service,
although they do have some material specific limitations. Applications are usually
limited to an impeller circumferential speed of 23 m/s (75 ft/s), corresponding to a
total head per pump of about 40 m (131 ft), although this can be increased through
the use of stiffer elastomers, at some expense to wear resistance. Chemical compo-
sition, temperature, and the presence of petroleum products in the slurry must be
considered. Impellers employing elastomers often have a higher NPSHR because of
the thicker impeller vane sections needed, and this condition may further limit their
use. Pump casings made entirely of elastomeric materials have insufficient strength
to withstand the pressure loads, so it is necessary to have an outer casing of some
sort, a configuration that is commonly called a double-wall design. For versatility,
some manufacturers make these designs for interchangeable wear parts of hard
metal, rubber, or urethane as the service warrants.
For large slurry pumps, the double-wall configuration is heavy, costly, and
difficult to maintain, and an alternative “tie-bolt” construction, using a hard metal
casing sandwiched between ductile metal outer plates, is sometimes used. The
tie-bolts provide a safety enclosure for the casing and lend flexibility to the
pressure-containing parts, making them more resistant to unplanned pressure surges
and therefore suited to use in long-distance pipelines that are subject to water
hammer. Figure 8.15 provides representative views of the various wet-end types.
While white iron and various elastomers remain the materials of choice for slurry
applications, other materials, such as ceramic inserts, welded overlays, and metal
matrix composites are coming into common use. Further information on wear and
wear-resistant materials is given in Chap. 13.
The data in Table 8.1 were collected during a pump performance test on clear water.
Calculate the pump efficiency, and then scale and plot the pump performance to a
constant speed of 435 rpm. Next, consider how the pump performance will change
when pumping slurry at the same speed.
To calculate pump efficiency, first use Eq. (8.2) to calculate the output power of
the pump. The efficiency is equal to the ratio of output-to-input power (Eq. 8.5).
Remember to use a consistent set of units.
To scale the flow (Q), head (H ), and power (Pinput) to the new speed of 435 rpm,
use the affinity scaling laws of Eqs. (8.6), (8.7), and (8.8). Remember that Q, H, and
P must be scaled together. Note that the tested speed varies from point to point in the
data, so a different speed ratio must be used when scaling each point. Since this is a
larger pump (judging by the flow rates) and the speed change is relatively small, the
adjustment to efficiency will be small and can be neglected. The results of the
calculations are shown in Table 8.2 and plotted in Fig. 8.16.
When running on slurry, the pump performance will change in several important
ways. The power input will increase in proportion to the density of the slurry. In
addition, some amount of head and efficiency will be lost due to the effect of the
solids. This solids effect may be negligible with fine settling slurries, but it can be
quite significant with slurries having a high density, coarse solids, or increased
viscosity. The solids effect should always be considered during slurry pump selec-
tion and will be explored in depth in the Chap. 9.
90 4500
80 4000
70 3500
Head H (m) and Efficiency h (%)
60 3000
50 2500
40 2000
Power PINPUT (kW)
30 1500
20 1000
10 500
0 0
0 2000 4000 6000 8000 10000 12000 14000
Flowrate Q (m3/hr)
Table 8.3 and Fig. 8.17 give the results of several NPSHR tests performed on clear
water in the test lab. Calculate the NPSHR3% and NPSHR5% for each test, and then
scale and plot the NPSHR curves to a constant speed of 350 rpm. Next, consider how
the NPSHR performance will change when pumping slurry at the same speed.
First, use the data provided at the tested pump speed of 285 rpm to estimate the
NPSHR3% and NPSHR5% per the method outlined in Sect. 8.3 and Fig. 8.11. Then,
scale the flow (Q) and NPSHR to the new speed of 350 rpm, using the affinity scaling
laws of Eqs. (8.6) and (8.7). NPSHR scales according to the same Eq. (8.7) as used
8.5 Case Studies 259
Table 8.3 Pump NPSHR NPSHR test 01 NPSHR test 02 NPSHR test 03
test data
Q = 1170 l/s Q = 1525 l/s Q = 1835 l/s
N = 285 rpm N = 285 rpm N = 285 rpm
H NPSHA H NPSHA H NPSHA
(m) (m) (m) (m) (m) (m)
17.87 18.50 16.68 19.50 15.52 19.90
17.90 10.60 16.68 16.10 15.55 13.10
17.92 6.00 16.70 10.30 15.62 7.10
17.92 3.60 16.72 3.90 15.54 4.10
17.90 2.80 16.71 2.90 15.45 3.30
17.74 2.10 16.60 2.30 15.50 2.40
17.50 1.70 16.35 1.95 15.05 2.00
17.33 1.55 16.18 1.70 14.90 1.95
16.98 1.35 16.05 1.60 14.74 1.90
16.44 1.25 15.85 1.55 14.48 1.80
18
Q = 1170 l/s
Head H (m) at 285 rpm
17
Q = 1525 l/s
16
Q = 1835 l/s
15
14
0 5 10 15 20
NPSHA (m)
for head. The tested and scaled NPSHR values are shown in Table 8.4 and plotted in
Fig. 8.18.
When running on slurry, all NPSHA values should be expressed in head of the
mixture being pumped. Regarding the solids effect on NPSHR, little data is available
since NPSHR testing on slurries is not common. For settling slurries, the effect is
generally observed to be smaller than the solids effect correction used for pump
head. There is one important exception: if the slurry mixture is significantly more
viscous than water, whether due to a viscous carrier fluid or the rheological effects of
a high solids concentration, NPSHR may be significantly affected. See Chap. 9 for
more information on the pump solids effect.
260 8 Centrifugal Slurry Pumps
3.5
NPSHR3%
3.0
NPSHR5%
2.5
at 350 rpm
NPSHR (m)
2.0
1.5
at 285 rpm
1.0
0.5
0.0
1000 1200 1400 1600 1800 2000 2200 2400
Flow rate Q (liters/sec)
Fig. 8.18 As tested and scaled 3% and 5% NPSHR curves
The pump from Case Study 8.1 is running at a fixed speed of 435 rpm on clear water.
As seen in Table 8.2, it produces 50.7 m of head at a flow rate of 8814 m3/h. The
operator wants to reduce this to 42 m at the same flow and speed by trimming the
impeller diameter. The current impeller vane diameter is 1500 mm. Find the required
diameter reduction, and plot the resulting performance curves for the trimmed
impeller.
As mentioned in Sect. 8.2, there is no exact formula to predict the performance of
a trimmed impeller. In this example, we use the method where flow is scaled in direct
proportion to the trimming ratio, and then we apply a correction factor to the
trimmed diameter to account for the expected error. An efficiency correction is
also applied.
With this method, the scaling equations mimic the effect of a speed ratio, as seen
below:
8.5 Case Studies 261
65
60
55
50
Full sized impeller
Head H (m)
45 at 435 rpm
40
Target duty point
Q = 8814 m3/hr
35
H = 42 m
30 Scaled for impeller trim
TD = 0.93
25
5000 7000 9000 11000 13000
Flowrate Q (m3/hr)
DTRIM
TD = ð8:18Þ
DFULL
Q2 = Q1 TD ð8:19Þ
H 2 = H 1 TD2 ð8:20Þ
P2 = P1 TD3 ð8:21Þ
The trimming ratio must be determined by an iterative approach, since both flow
rate and head scale together. Because of this, the point on the head curve that scales
to the new target performance will come from a flow rate higher than the target flow
rate. The resulting trimming ratio for our example is 0.93, as seen in Fig. 8.19.
Based on this trimming ratio, the trimmed diameter should be 1395 mm. How-
ever, this scaling method will underestimate the impeller diameter. One type of
empirical correction factor is to increase the diameter by some percentage of the
difference between the full-sized and trimmed diameters, based on experience with
similar designs. This may be in the range of 20% for slurry pumps, as seen in
Eq. (8.22). Applying this correction results in a final trim diameter of 1416 mm:
The trimmed impeller may have a lower efficiency than the full-sized version. If
the trim is small (less than 5%), this fact may be neglected; larger trims, however,
will require some correction. The correction will tend to increase with increasing
value of the TD ratio and increasing specific speed. Eq. (8.23) gives an empirical
relationship that can be used as an approximation for slurry pumps of specific speed
NS < 45:
262 8 Centrifugal Slurry Pumps
90 4500
80 4000
Head H (m) and Efficiency h (%)
70 3500
60 3000
50 2500
40 2000
20 1000
10 500
0 0
0 2000 4000 6000 8000 10000 12000 14000
Flowrate Q (m3/hr)
0:022N s
η2 = η1 ð - 1:2TD2 þ 2:5TD - 0:3Þ ð8:23Þ
Calculating NS from the BEP point in Table 8.2 gives NS = 35.8, providing an
efficiency correction factor of 0.990 at the TD ratio of 0.93. Figure 8.20 shows the
final estimated performance for the trimmed impeller compared with the full-sized
performance.
The above scaling relations for trimming have limitations and should not be used
for TD < 0.8. Special care must also be taken with slurry pumps of specific speed
NS > 45, as these are more sensitive to trimming and the resulting reduction in the
vane overlap. When in doubt, consult the pump supplier for specific recommenda-
tions, or plan a performance test with the trimmed diameter. It is common practice to
test multiple trimmed diameters for standardized water pumps, but it is less common
for slurry pumps due to the materials of construction and frequent application of
variable speed drivers.
References
H.H. Anderson, Prediction of head, quantity and efficiency in pumps – The area ratio principle, in
Performances Prediction of Centrifugal Pumps and Compressors, ed. by S. Gopalakrishnan,
(ASME, New York, 1980)
ANSI/HI 12.1–12.6, American National Standard for Rotodynamic Centrifugal Slurry Pumps
(Hydraulic Institute, Parsippany, 2021)
ANSI/HI 14.6, American National Standard for Rotodynamic Pumps for Hydraulic Performance
Acceptance Tests, Appendix K (Hydraulic Institute, Parsippany, 2022)
References 263
ANSI/HI 20.3, Hydraulic Institute Program Guideline for Rotodynamic Pump Efficiency Predic-
tion (Hydraulic Institute, Parsippany, 2020)
M.R. Carstens, T.W. Hagler, Water hammer resulting from cavitating pumps. J. Hydraul. Div.
ASCE 90, 161–184 (1964)
P. Cooper, Centrifugal pump theory, in Pump Handbook, ed. by I.J. Karassik, 3rd edn., (McGraw-
Hill, New York, 2001), pp. 2.1–2.96
L. Euler, Théorie plus complette des machines qui sont mises en mouvement par la réaction de
l’eau. Societas Scientiarum Naturalium Helveticæ 10, 227–295 (1756)
I.J. Karassik, J.P. Messina, P. Cooper, C.C. Heald, Pump Handbook, 4th edn. (McGraw Hill,
New York, 2008)
C.P. Kittredge, P. Cooper, W.H. Fraser, Centrifugal pumps: Hydraulic performance and behavior,
in Pump Handbook, ed. by I.J. Karassik, 4th edn., (McGraw-Hill, New York, 2008),
pp. 2.132–2.136
S. Lazarkiewicz, A.T. Troskolanski, Impeller Pumps (Pergamon, Oxford, 1965)
A.J. Stepanoff, Centrifugal and Axial Flow Pumps, 2nd edn. (Wiley, New York, 1957)
F.M. White, Fluid Mechanics, 8th edn. (McGraw Hill, New York, 2016)
R.C. Worster, The flow in volutes and its effect on centrifugal pump performance. Proc. Inst. Mech.
Eng. 177, 843–875 (1963)
H. Yoda, K. Uranishi, M. Oshima, T. Sakurai, D. Torii, Hydraulic efficiency conversion from a
model to prototype pump based on effects of reynolds number and surface roughness. Presented
at the ASME Fluids Engineering Division Summer Meeting FEDSM, Washington, DC, 2016
Chapter 9
Effect of Solids on Pump Performance
The presence of solid particles in the flow tends to produce adverse effects on pump
performance, and detailed information on these effects is needed to achieve reliable
and energy-efficient operation. Since a centrifugal pump always operates at the
intersection of the pipeline and pump head characteristics, degradation of the
pump characteristic due to slurry solids can alter the operating point away from
the intended design. A variable speed drive may be used to adjust for these effects
and restore the desired operating condition; however, even a small deviation from a
predetermined average pump speed can be costly. The deviation of a few percent in
speed may also markedly influence pump wear in severe applications. Furthermore,
the pipeline characteristic for settling slurry flow very often displays a minimum
(as shown in Chaps. 4 and 5 ), followed by a slow rise with increasing flow rate. The
full implications of this type of intercept will be examined in Chap. 10 , but it is
important to note here that even a small diminution in pump head can produce a
disproportionately large reduction in flow rate.
The analysis of solids effect on pump performance is not yet as well-developed as
that detailed in earlier chapters for the effect of solids on resistance to flow in
pipelines, but similar lines of reasoning can be employed, as will be shown. There
are two related phenomena, however, that have been studied in detail in connection
with pumping simple fluids: the effects of varying density and the effects of varying
(Newtonian) viscosity.
If only the fluid density is changed while the viscosity remains constant, the
affinity laws provide a direct method of evaluating the results (Sect. 8.1.2). At a
given flow rate (for constant pump size and rotational speed), the internal flow
characteristics within the pump are unchanged. Hence the efficiency is unaffected,
and so is the head, when measured in terms of height of the fluid being pumped. The
relative density of the fluid (to that of standard water) is Sf, and the resulting power
requirement is Sf times that needed to pump water at an equal pump head and flow
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 265
R. Visintainer et al., Slurry Transport Using Centrifugal Pumps,
https://doi.org/10.1007/978-3-031-25440-6_9
266 9 Effect of Solids on Pump Performance
η
Water density
Higher density
Flow rate
rate. For pumps of a given size and rotational speed, it can be readily shown that the
flow rate at the best efficiency point is not influenced by the density change.
Likewise, the density change effect does not depend on pump size—provided, of
course, that the comparison is made within a set of geometrically similar (homolo-
gous) machines.
As a slurry is usually denser than water, it appears attractive to use the behavior of
a denser fluid as a model of slurry behavior, though it may only be an approximate
one. This approach corresponds to the equivalent liquid model, considered in
Chap. 2 as a first approximation for slurry flow in pipes. Figure 9.1 represents the
results of applying the equivalent liquid model to a centrifugal pump.
Note that the head is expressed in height of the fluid being pumped, and the head
produced when pumping the higher-density equivalent liquid is numerically the
same as that produced when pumping water at all flow rates.
As mentioned above, the effects of a change in fluid viscosity are also well-
known. If the density of the fluid being pumped is kept constant while its viscosity is
increased, then both the head and efficiency decrease, while the power needed to
maintain a constant flow rate increases. This behavior is shown in Fig. 9.2.
Note that the effects of viscosity are more pronounced at higher flow rates, and as
a result, the flow rate at the best efficiency point tends to become smaller as viscosity
is increased. This shift may not always be significant, but another influence, which
cannot be represented on the curves for a single pump, is of greater importance. This
is the size or scale effect: an increase in viscosity has a greater influence on small
pumps than on large ones.
9.1 Introduction and Definitions 267
Water viscosity
Higher viscosity
Flow rate
H Hw
Hm
hw
hm
h
Water
Settling Slurry
Flow rate
hm
Q Q0
Hm
Hw = ð9:1Þ
HR
The pump must therefore operate at a higher rotary speed that corresponds to the
increased head requirement represented by Hw. The derated value of the slurry
efficiency (ηm) is then calculated by multiplying ER with the water efficiency
obtained from this speed.
The power output of a pump is determined by the product of the slurry or water
pressure (ρmg Hm or ρw gHw, respectively) and the flow rate (Q), divided by the
corresponding slurry or water efficiency:
ρm gH m Q
Pm = ð9:2Þ
ηm
ρw gH w Q
Pw = ð9:3Þ
ηw
The ratio of the power requirements for slurry (Pm) and water (Pw) from
Eqs. (9.2) and (9.3) can be expressed as follows, together with the definitions in
Fig. 9.3:
270 9 Effect of Solids on Pump Performance
Pm HR 1 - RH
= Sm = Sm ð9:4Þ
Pw ER 1 - Rη
When the reductions in head and efficiency are equal (i.e., HR = ER or RH = Rη),
then the power consumption when pumping slurry increases directly with the
relative density of the slurry:
Pm = Sm Pw ð9:5Þ
In situations when HR > ER or RH < Rη, then Eq. (9.4) corresponds to:
Pm > Sm Pw ð9:6Þ
In this case, the power consumption increases more than expressed by the relative
slurry density. Examples of results with HR > ER for metal pumps with thick
hydraulic sections for severe operating duties, and for rubber-lined pumps in general,
are shown in Sect. 9.2.1.
The solids properties and individual pump design factors have more varied and
less predictable effects on reductions in efficiency than they do for those in head, and
modeling of the efficiency reduction is therefore more complex and less accurate.
Recent large-scale loop-test findings with metal slurry pumps show that HR corre-
lates reasonably well, but not precisely, to the assumption that HR = ER (Visintainer
et al. 2017, 2021, 2022). However, in settling slurries where the fractional solids
concentration is high (Cv > 0.35), or the solids are very coarse, the efficiency
reduction may be considerably greater, and a more precise analysis may be desired.
Small- and medium-sized metal pumps for severe operating and wear duty with thick
material sections may also exhibit HR > ER (i.e., Pm > Sm Pw), for concentrations as
low as Cv = 0.2. The same approximate limit seems to hold for rubber-lined pumps.
For highly concentrated fine-particle slurries (i.e., those with significant viscous
and/or non-Newtonian effects), the efficiency derating is normally greater, as cov-
ered in Sect. 9.3.
Rather than attempting to calculate the efficiency solids effect, it is often assumed
to be similar to the head effect, and a margin of 15–25% is then placed on the
installed driver power to account for any uncertainty. Ultimately, the head effect is
most important for system design, while efficiency is largely concerned with the
sizing of the pump driver.
Although separate from the solids effect, additional deratings of 10–15% should
also be anticipated for highly abrasive slurries due to the wear of the pumping
components over time.
9.1 Introduction and Definitions 271
The derating procedure is now demonstrated in detail for a metal pump with impeller
diameter of D = 0.4 m and discharge diameter = 0.1 m, in an example based on
loop-test results with a heavy ore product (Visintainer et al. 2021).
The solids specific gravity is 4.75, and the particle size distribution is broad, with
12% of the particles less than 0.04 mm, a d50 = 3 mm, and dmax of about 10 mm.
Operation is at Cv = 0.234. Loop test results showed an average solids effect on head
of RH = 0.27 (i.e., HR = 0.73) and a similar value for the average solids effect on
efficiency, ER = HR.
A slurry head of 22.5 m is desired at a flow rate of 128 m3/h (0.0356 m3/s). To
produce the required head on slurry, the pump must be capable of producing a water
head Hw = Hm/HR (Eq. (9.1)) at the same flow rate, so that Hw = 22.5/0.73 = 30.8 m
at 128 m3/h. Assuming a pump head-capacity characteristic as seen in Fig. 9.5, the
rotary speed must be increased from about 1000–1146 rpm. The water efficiency
(ηw ) at this speed for the 128 m3/h flow rate is about 67%.
The slurry density and efficiency are calculated as follows, assuming a water
density ρw = 1000 kg/m3, and applying the assumption that ER = HR:
Fig. 9.5 Pump head-capacity characteristics used to demonstrate the derating procedure
272 9 Effect of Solids on Pump Performance
ρm gH m Q
Pm = = 1878: 9:81: 22:5: 0:0356=0:49 = 30116 W or 30:12 kW
ηm
Note that if ER were known to be smaller than HR, then the power requirement
would be correspondingly larger, according to Eq. (9.4). Also note that the measured
deratings are based on laboratory test-loop data; therefore, bear in mind that the
reductions in head, and especially in efficiency, may be somewhat larger in practice,
since the closed-loop nature of laboratory testing often results in more solids
degradation than is seen in the typical “once-through” industrial application.
Early investigations into the pump solids effect were reviewed by Stepanoff (1965),
who pointed out the dependence on particle size (d), solids concentration by volume
(Cv), and solids density ratio (Ss). Head reduction curves for narrowly graded
products were presented by McElvain (1974) in terms of a graph with Cv, Ss, and
d50, partly based on findings later presented by Burgess and Reizes (1976) and Cave
(1976). Loop-test data for various industrial products confirmed approximately the
curves for rubber-lined and metal pumps with impeller diameters of 0.28 to 0.43 m
(Fairbanks 1941; Vocadlo et al. 1974; Burgess and Reizes 1976; Sellgren 1979).
Czarnota et al. (1996) noted that it was difficult to compare test results over time,
since developments in slurry pump hydraulic design resulted in head and efficiency
reductions (RH and Rη) in the mid-1990s, which were about half of those reported
during the 1970s.
The dependence of the pump size was not addressed in these early investigations,
especially in the abovementioned studies. Sellgren and Addie (1989) presented loop-
test results on how the head derating effect decreased with increasing pump size for
narrowly graded sands and gravel in pumps with impeller diameters from 0.25 to
1.2 m. In the previous edition of this book (Wilson et al. (2006), HR and RH were
presented in a generalized diagram for different pump sizes (impeller diameters)
versus the particle size (d50) for narrowly graded particles. Corrections for Cv, Ss,
and the fine-particle content of less than 75 μm were then adopted concurrently. This
method was later adopted by the Hydraulic Institute centrifugal slurry pump standard
ANSI/HI 12.1–12.6 in its original and subsequent editions from 2005 through 2021.
Engin and Gur (2003) presented an extensive study covering a large variety of
metal and rubber-lined pump test results available in the literature. Both the data
used and the resulting correlations were mainly limited to small pumps with impeller
diameters from 0.21 to 0.45 m. Their resulting relationship for the reduction in head
(RH) was expressed as 2.71(Cw. (Ss-1)0.65) . (d/D)0.313, where Cw is the fractional
solids concentration by mass = Cv Ss/Sm, and d is related to a weighted particle size,
or approximated by d50. Their formula tends to overestimate RH, possibly due, in
part, to an overrepresentation of older literature data in their analyses.
9.2 Settling Slurries with Newtonian Liquids 273
Visintainer et al. (2017, 2021, 2022) performed a series of tests on a wide range of
settling slurries of various solids concentrations, sizes, and densities, and using
pumps of various sizes from 310 mm to 1435 mm impeller diameters. From this
data set, a 4-Component Model (4CM) for pump solids effect was developed, similar
in concept to the pipe friction model presented in Chap. 5.
The following text describes the ANSI/HI and 4CM models in greater detail, after
first describing the individual effects of the main variables influencing the pump
solids effect: solids concentration, solids density, pump size, and particle size
distribution.
The comprehensive test results by Visintainer et al. referred to above, which used
standard (or “contemporary design”) metal pumps, confirmed a linear dependence
between concentration and head deration, and an approximate equivalence between
ER and HR at delivered concentrations up to and somewhat above Cv = 0.3.
However, the magnitude of the increase with concentration was very dependent on
the particle size distribution, with coarse solids containing few fine particles having
the strongest effect, while much weaker derates resulted from broadly sized distri-
butions, even at very high concentrations. There was a limit to this linear depen-
dence; at a certain high Cv, depending on the particle size distribution, the solids
effect began to increase at a higher rate.
As stated previously, for extremely high values of the concentration, the equiv-
alence of the head reduction factor and the efficiency reduction factor no longer
holds, and the fractional reduction in efficiency is more than that in head. As a result,
the power consumption must be larger than that given by Eq. (9.5) (i.e., Pm > SmPw).
Figure 9.6 shows the detailed development of the deratings from 0.35 to 0.45 Cv for
a narrowly distributed sand with d50 = 0.35 mm and an all-metal pump with impeller
diameter of 0.63 m. Ni et al. (1999) presented similar results for concentrations up to
about 0.35 when pumping a narrowly graded sand with a d50 of 370 μm in an
all-metal, three-vane pump with an impeller diameter of 0.4 m.
Centrifugal slurry pumps can be used for very high solids concentrations. Exper-
imental results from the GIW Hydraulic Laboratory with coarse sand in 0.25 and
0.30 m diameter pipe loop systems achieved an upper limit of solids concentration
by volume of about 0.49. This limit was not set by the pump, but by dramatically
varying pipeline friction losses with unstable operating conditions (Addie and
Hammer, 1993). The average and maximum particle sizes were about 0.6 and
10.0 mm, respectively, and the pump had an impeller diameter of 1.1 m. Large-
scale dredging field data by Vercruijsse and Corveleyn (2002) with a 2.6 m impeller
diameter pump covered Cv values up to nearly 0.5.
274 9 Effect of Solids on Pump Performance
In many cases, these slurries can be treated as settling slurries in both the pump
and pipeline. However, at delivered volumetric concentrations exceeding 0.45, and
especially in the case of intermediate particle size distribution slurries with average
sizes of 20–100 μm and maximum size of up to about 1 mm (representing, for
example, thickened tailings), the slurry begins behaving more like a viscous single-
phase fluid. In these cases, it may be better modeled using the methods developed for
non-Newtonian rheologies considered in Sect. 9.3. This also obviously applies to
any slurry whose carrier fluid rheology is significantly non-Newtonian, due, for
example, to the presence of significant amounts of clay or other fine and rheologi-
cally active solids.
Lined slurry pumps often have thicker hydraulic sections, especially in the impeller
vanes and shrouds, as compared to similar-sized, solid metal designs, and they may
see a deviation between the head and efficiency derates occurring at lower concen-
trations. Sellgren and Vappling (1986) conducted tests for two different rubber-lined
pumps with impeller diameters of about 0.45 m pumping industrial mineral solids
with d50 of 0.35 and 0.1 mm. Their results showed a sharp drop in ER starting at a Cv
of about 0.2, and HR/ER values of 1.6 and 1.3, respectively, at Cv = 0.35, meaning
that Pm > SmPw. A largely similar trend was seen 20 years later for results at the
GIW Hydraulic Laboratory with a three-vane, rubber-lined pump with an impeller
diameter of 0.4 m handling a narrowly graded sand product with d50 = 0.6 mm, as
shown in Fig. 9.7.
9.2 Settling Slurries with Newtonian Liquids 275
HR,ER
0.6
HR
0.5
0.4 ER
0.3
0 0.1 0.2 0.3 0.4
Solids concentration by volume
A number of studies have been made on the effect of solids density on pump
performance. McElvain (1974) proposed that the head- and efficiency-reduction
factors, RH and Rη, are directly proportional to the delivered concentration by
volume and expressed graphically the effect of Ss and average particle size (d50),
mainly based on experimental data for sands and heavy ores. Kazim et al. (1997) and
Engin and Gur (2003) expressed the solids effect in terms of Cw(Ss-1)λ where λ is a
constant less than 1. Engin and Gur (2003) analyzed heavy ore and sand results by
Burgess and Reizes (1976) and diverse solids density test results available in the
literature with a solids density ratio other than 2.65. They reported uncertainties with
variation in λ from 0.5 to 1. Wilson et al. (2006) related the slurry density as a leading
parameter, of which Cv(Ss - 1) should be a natural expression; however, following
the findings by Engin and Gur (2003), Cv((Ss - 1)/Ssref)0.65 was chosen in the
diagram method. This was later incorporated into ANSI/HI 12.1–12.6, where Ssref is
a reference value for sand and gravel with Ss = 2.65, i.e., Ssref = 2.65 – 1 = 1.65.
In the above relations, a fluid density of Sf = 1 is assumed. In extensive testing by
Visintainer et al. (2017, 2021, 2022) to support the development of the 4CM (which
included an iron ore product with Ss = 4.75), a formulation of Cv((Ss - Sf)/Ssref)λ
was used, with Sf taken as the density ratio of the fluid plus the solids included in all
finer components as each coarser component was evaluated. In this case, a value of λ
= 1 was determined to provide the best fit to the data, indicating a linear effect of
solids density similar to that of concentration, when considering the incremental
contribution of each narrowly graded, particle size fraction.
Much of the available solids effect data are for smaller pumps (<1 m impeller
diameter), and in general, these show a weak effect of pump size. This was observed
276 9 Effect of Solids on Pump Performance
In the chart developed for the previous edition of this book and incorporated in the
ANSI/HI 12.1–12.6 standard (which is described in more detail below) the particle
size dependence is given by the function:
α
d
FnðdÞ = 50 ð9:7Þ
d ref
where dref = 0.001 m (i.e., 1 mm) and α = 0.4 . (d50/dref)-0.25 with d50 expressed in
the same units as dref. In general, it tends to under-predict losses for narrowly graded
coarse solids and over-predict losses for slurries with a broad particle size distribu-
tion where the ratio of d85/d50 is large.
The data collected by Visintainer et al. (2017, 2021, 2022) associated with the
development of the 4CM included more than 80 tests covering a range of particle
sizes from < 40 μm to 20 mm, with d50 particle sizes from < 40 μm to 10 mm, and
delivered solids concentrations from 0.04 to 0.38 by volume. Particle size distribu-
tions varied from very narrow to very broad, with d85/d50 ratios from 1.3 to 30 and
beyond. Solids included silica-based rock and high-density iron ore. To limit particle
degradation, solids were loaded into the loop system to minimize recirculation and
extended run times were avoided. To ensure accurate characterization of the particle
sizes being pumped, slurry samples were taken using a device that cut through the
9.2 Settling Slurries with Newtonian Liquids 277
entire slurry flow stream. By accounting for the contribution of each of the four size
ranges (or components) separately, rather than depending on a single d50, the particle
size dependence exponent of Eq. (9.7) was simplified to a constant value of α = 0.4.
The literature contains few other experimental derating results for very coarse
particle slurries in both small and large pumps, and particularly with industrially
crushed products; however, the dependence expressed in Eq. (9.7) with α = 0.4 does
support older test loop data. This includes data with dmax in excess of 20 mm for
pumps with diameters in the range of 0.8 to 1 m with gravel and rock by Mez (1984),
Sellgren and Addie (1989), Shook et al. (1995), and for iron ore with Ss = 4.35 by
Sellgren et al. (1997).
9.2.5 Modeling
This section presents two detailed models for general use with settling slurries. Both
have already been introduced previously with regard to some of the particular effects
of key variables.
The mono-sized method from the previous edition of this book is first described in
equation form. This method is relatively easy to apply and produces reasonable
results of proven use for many common industrial slurries. It is widely known
through its incorporation into the Hydraulic Institute Centrifugal Slurry Pump
Standard ANSI/HI 12.1–12.6 (2021).
The 4-Component Model (4CM), based on the pipe friction model of the same
name introduced in Chap. 5, partitions the slurry solids into four fractions, each of
which contributes to the total pump solids effect based on its average size and
concentration within the total slurry. This model necessarily involves more calcula-
tion and a fuller description of the particle-size distribution, but it also provides
additional accuracy, especially for very coarse or very broad particle-size distribu-
tions, as well as for unusually sized distributions not following a mathematically
“normal” curve, such as artificial mixtures of coarse and fine slurries sometimes used
in tailings placement. For many typical industrial slurries, the two models will
produce similar results. Where they differ, the 4CM formulation generally provides
greater accuracy.
The generalized diagram procedure for settling slurries incorporated into the
Hydraulic Institute Standard for Centrifugal Slurry Pumps ANSI/HI12.1–12.6
(2021) is based on the following relationships from Wilson et al. (2006). RH is
given in terms of pump impeller diameter (D) and solid size (d50), with corrections
for concentration (Cv), relative density of solids (Ss), and fraction of particles less
than 75 μm (X0):
278 9 Effect of Solids on Pump Performance
- 0:9 α 0:65
D d 50 Cv Ss - 1
RH = k ð1 - X 0 Þ2 ð9:8Þ
Dref dref C vref Ssref
where RH is given as a fraction and the constant is related to the pump size.
Reference values are as follows: Dref = 1.11 m; dref = 0.001 m (i.e., 1 mm);
Cvref = 0.15, and Ssref = 1.65. Both D and d50 must be expressed in the same
units as Dref and dref, respectively.
The size effect factor (k) = (0.0187+0.0521D) with D in meters for
0.41 < D < 0.89. The factor is fixed at 0.04 for impeller diameters smaller than
0.41 m, and at 0.065 for diameters larger than 0.89 m. The dependence of the
derating on pump size is expressed both by this factor and the -0.9 exponent. If
the influence of the factor (k) were included in the size dependence exponent while
holding the factor itself = 0.065, the exponent would become a variable ranging
from about -0.5 for smaller impellers (0.4 to 0.8 m diameter) up to -0.9 for larger
pumps.
Sellgren and Addie (1989) adopted the standard particle drag coefficient (CD) to
characterize the particle-related resistance within the two-phase flow field where
particles experience abrupt directional changes due to the highly turbulent flow within
the centrifugal pump. The particle size dependence was related to the inverse root of
CD, which is proportional to the particle terminal settling velocity. In order to cover a
span that includes large particles, the size dependence in Eq. (9.8) was given by the
empirical expression in Eq. (9.7), namely (d50/dref)α where α = 0.4(d50/dref)-0.25, with
d50 and dref expressed in the same units. For details, see Sellgren et al. (2017).
The influence of particles smaller than 75 μm (X0) is schematically included
through (1-X0)2. This approximates the reduced settling velocity of larger particles
due to the fine particle content in terms of the density difference and the
corresponding equivalent liquid behavior.
Wilson et al. (2006), Shook et al. (1995), and others have pointed out the similarities
of pump solids effect mechanisms to the influence of leading parameters on slurry
friction losses in pipelines. The observation of a velocity dependence also supports
the presumption that there is only one set of physical mechanisms involved, and thus
it is reasonable to consider a link between the models used for solids effect in pumps,
and those used in pipelines. Figure 5.14 shows a favorable effect on pipeline friction
losses for a settling slurry of broadly sized distribution (d50 = 0.2 mm, dmax = 10 mm)
and high concentration by volume of 0.379. The comparatively small losses are
related to the wide distribution of d85/d50 (in this case over 10) and on how the
friction losses are flattening out for increasing solids concentrations above
0.15–0.20, as shown in Fig. 5.16. Evaluation of the solids effect on head for the
0.81 m impeller pump used in this case experienced a head reduction (RH) of about
0.075. In contrast to this, a much higher RH of 0.20 was observed for coarse particles
9.2 Settling Slurries with Newtonian Liquids 279
alone, having a d50 of 7.4 mm and dmax of 12.5 mm, but substantially lower Cv of 0.2
(i.e., only about half the solids content). Test results by Visintainer et al. (2022) with
a large pump (impeller diameter of 1.44 m) and particles of up to 22 mm amplified
the differences seen in the example above. Those results showed strong head loss
effects for narrowly graded, coarse-particle slurries at moderate concentrations, and
comparatively small effects for mixtures with a broad particle size distribution at
higher solids concentrations.
For very coarse, narrowly graded rock particles, Shook et al. (1995) discussed
how the transition mechanisms via the fluid between the impeller and the casing
must be coupled to significant relative velocity dissipation. In an analysis related to
pump design, Roco et al. (1986) looked at water and slurry losses in a 0.81 m
impeller pump. Results indicated that one-third of the energy losses related to slurry
flow could be associated with acceleration transfers in the impeller-volute region for
a sand with an average particle size of 0.25 mm and moderate Cv of 0.18.
Recognizing that the particle size dependency of pump solids effect mechanisms
is similar to leading parameters of slurry friction losses in pipelines, suggests the use
of a multi-component formulation similar to the 4CM pipe flow model of Sect. 5.5.1,
where a calculation and summation of individual pump performance losses is made
according to the relative volumetric concentration of each solids component (i.e., the
carrier fluid (Xf), pseudo-homogeneous (Xp), heterogeneous (Xh), and fully stratified
(Xs) solids fractions). In transferring this formulation to a pump solids effect model,
similar categories of particle size are used where each component is represented by
its average particle size and provides its individual contribution to head loss. The
component contributions are then summed up to provide the total head loss (RH).
Based on the similarity with pipeline friction losses, this approach ultimately led
to the use of the empirical parameters A″, B″, and C″, as used in the pipeline model.
These parameters are related to a pipe size, which in the pump solids effect model is
represented by the pump discharge diameter (DD). They account for the action of
finer particles in reducing the derate contribution of coarser particles, including the
dependence on flow rate, based on the limiting pipeline velocities of deposition and
pseudo-homogeneous flow.
For adaptation to the 4CM, the mono-size basis of Eq. (9.8) was updated to recent
findings presented in the preceding sections according to the following generic
format:
-D0:5D α
D ref d 50,x C v,x Ss - 1
RH,x = 0:08 ð9:9Þ
Dref d ref Cvref Ssref
where the “x” subscripts refer to the individual particle size fractions associated with
the 4-component model. The exponent of particle size dependence α = 0.4, as
described in Sect. 9.2.4. The reference particle size dref = 0.001 m (i.e., 1 mm).
The impeller diameter dependence is related to the exponent -0.5(D/Dref) with
Dref = 1 m. Both D and d50 must be expressed in the same units as Dref and dref,
respectively. The concentration and solids density reference values (Cvref and Ssref)
280 9 Effect of Solids on Pump Performance
1
0.95
0.9
0.85
0.8
HR
0.75
0.7
4CM
0.65
ANSI
0.6
0.55
0.5
0.01 0.1 1 10 100
Particle size (mm)
Fig. 9.8 HR versus particle size (d ) for the mono-size and 4-Component Model bases. Cv = 0.15,
Ss = 2.65, impeller diameter = 0.8 m
remain at 0.15 and 1.65, respectively. The exponent dependence for the specific
gravity term is set = 1, as discussed in Sect. 9.2.2.
Figure 9.8 compares the mono-sized and 4CM models at Cv = 0.15, Ss = 2.65,
and a pump impeller diameter of 0.8 m. In the case of the 4CM, a typical particle size
distribution with a mathematically “normal” distribution and having d85 = 2.3d50 is
assumed. The drop in the 4-Component curve at particle sizes above 1 mm expresses
the larger weight given to these coarse particles when treated as one part in a sum of
components for a particle size distribution. Note that this stronger effect of the larger
solids will, however, be modified by the presence of smaller solids through the
empirical parameters A″, B″, and C″, through which the strong effect of the particle
size distribution becomes manifest, with B″ and C″ also expressing the velocity
dependence of the losses relative to the deposition and pseudo-homogeneous
velocities.
The 4CM for the centrifugal pump solids effect is now described in detail. The
particles in the slurry are first partitioned by size into four volume fractions or
components, similar to the pipeline friction model described in Sect. 5.5.1. As
with the pipeline model, grains finer than 40(2.65/Ss) μm compose the Xf fraction
and are considered to be part of the carrier fluid, and grains larger than 40(2.65/Ss)
μm and smaller than 200 μm contribute to the pseudo-homogeneous fraction (Xp).
The heterogeneous fraction (Xh) spans particle diameters between 200 μm and 1.5%
of the pump discharge diameter (i.e., 0.015DD), rather than the pipeline diameter.
The pump discharge diameter is also used in determining the stationary deposition
and pseudo-homogeneous limiting velocities, which will be needed to calculate the
empirical parameters B″ and C″. All grains larger than 0.015DD comprise the fully
stratified solids component (Xs). The weight averaged sizes of the three larger
components (d50p, d50h, and d50s) are also needed. These fractions and particle
9.2 Settling Slurries with Newtonian Liquids 281
Fig. 9.9 Solids fractions for model components delineated on the particle size distribution curve.
sizes are shown graphically in Fig. 9.9, and as with the pipeline model, the fractions
are defined relative to the total volume of particles, so that:
Xf þ Xp þ Xh þ Xs = 1 ð9:10Þ
There are some important differences between the pipeline friction and pump
solids effect models. As seen above, the selection of the pipe diameter for determin-
ing the limiting velocities and the Xh-Xs particle size boundary is based on the pump
discharge diameter instead of the pipeline (which may differ), and the average
particle sizes (by weight) must be defined for all three larger fractions. There is
also a difference in how the contributions of the components are combined. In
principle, the solids effect derates for the four components can be combined in a
simple sum to provide the total pump solids effect, as shown in Eq. (9.11):
However, while an increase in the viscosity of the carrier fluid (Xf) component
will increase the viscous losses contributed by that component, it will also decrease
the losses contributed by the larger components, as it better supports these particles,
limiting losses due to the inertial effects of particle acceleration. In the extreme cases
of very high or very low viscosity, the Xf term will dominate or become negligible
(respectively), and the simple sum is not much affected by the above-described
interaction. However, if the losses due to the carrier fluid viscosity are of a similar
order of magnitude to the three other components, a simple sum will tend to
overestimate the actual losses. In these cases, Visintainer et al. (2021) proposed a
practical approach for viscous settling slurries through the calculation of a “root sum
of squares” weighted average between the pure viscous effects of the carrier (RH,f ),
282 9 Effect of Solids on Pump Performance
according to ANSI/HI 9.6.7 (2021) and the pure solids effect derating of the
remaining fractions considered within a low viscosity carrier (e.g., water):
h 2 i0:5
RH = R2Hf þ RHp þ RHh þ RHs ð9:12Þ
Note that the deration provided by ANSI/HI 9.6.7 for Newtonian viscous liquids
varies with flow rate and is not a fixed value (see Fig. 9.2). Therefore, the total RH
deration relationship in (Eq. 9.12) must be evaluated on a point-by-point basis along
the pump curve. Fortunately, in many slurry applications, the RHf contribution will
be zero, and the formulation simplifies to a direct sum of the three larger component
contributions, as shown in Eq. (9.13).
The approach for calculating the contribution of each fraction is described in
detail below. In general, the carrier fluid (Xf) fraction is governed by viscosity, while
the other components follow the formulation of Eq. (9.9), each depending on the
average particle size of that fraction (i.e., d50p, d50h, or d50s) and a supporting mixture
density that includes the previous, smaller particle fractions (i.e., Sf, Sfp, or Sfph).
The contribution of the carrier fluid (RH,f) is considered to be entirely due to its
viscosity and may be determined by a standard method for viscosity corrections for
viscous Newtonian liquids, such as the ANSI/HI 9.6.7 (2021).
For many cases of practical interest, the carrier fluid is water, and the contribution
of the fine particles to the carrier fluid viscosity is small. In these cases, the effect of
fine solids on the dynamic viscosity of the carrier fluid may be determined by a
correction factor such as that proposed by Gillies et al. (1999) and given in
Eq. (5.15). However, in many cases, the resulting viscosity will be less than 0.02
Pas (20 cP) and the corresponding correction to the pump performance will be
negligible, in which case it may be assumed that RH,f = 0. For example, with an Xf of
0.20 in a slurry with a high Cv of 0.4, the corresponding increase in viscosity ratio by
Eq. (5.15) is only 1.6 (i.e., a viscosity of about 0.0016 Pas).
If the viscosity can be neglected, then RH,f = 0 and Eq. (9.12) reduces to:
X f C v ð Ss - S l Þ
Sf = Sl þ ð5:14Þ
1 - C v ð1 - X f Þ
A00 = 1 - X f þ 0:5X p ð5:21Þ
where Sf is the carrier fluid specific gravity and A″ is an empirical parameter from the
4-Component pipeline friction model expressing the friction reduction effects of
finer particles, repeated here from Chap. 5 using the default constants.
-D0:5D 0:4
00 D ref d 50h Ss - Sfp X h Cv
RH,h = C 0:08 ð9:15Þ
Dref d50ref 1:65 0:15
X f þ X p C v ð Ss - Sl Þ
Sfp = Sl þ ð5:18Þ
1 - Cv 1 - X f - X p
0:5
V 100,s - V m
C 00 = 1 - X f þ 0:5 X p ð5:28Þ
V 100,s - V sm,h
The reduction factor C″ is again an empirical parameter from the pipeline friction
model expressing the friction reduction effects of finer particles, (again with default
constants). It results in an increasing solids effect on head for increasing flow rates,
as determined by the velocity (Vm) in the pump discharge with diameter (DD). C″
varies from a minimum value at the deposition limit velocity for the d50h-sized
particle (Vsm,h), to unity at, or above, the velocity at full suspension for a ds-sized
particle (V100,s). Therefore, C″ = 1 for any Vm > V100,s. An estimated value for V100,s
may be determined according to the classical relationship by Newitt et al. (1955) for
the beginning of full suspension flow, here applied for ds = 0.015DD:
284 9 Effect of Solids on Pump Performance
p ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
V 100,s = 3
1800 g DD vt,s ð9:16Þ
where vt,s is the terminal settling velocity of the ds-sized particle. Note that for pump
discharge diameters DD > 0.075 m (75 mm), the value of ds corresponds to particle
sizes greater than about 0.0011 m (1.1 mm). The settling of a particle of this and
larger sizes can be approximately expressed by the relationship for turbulent settling
of a single particle (Eq. 2.57) which is independent of the liquid viscosity. It is here
empirically modified for settling of irregularly shaped sand and gravel (ξ ≈ 0.545 for
K = 0.26) following results by Haider and Levenspiel (1989):
Note that all velocities in Eqs. (9.16) and (9.17) must be expressed in m/s and all
diameters in meters. Vsm,h is the deposition limit velocity which can be obtained for
the d50h size from the nomogram in Fig. 4.8 or by Eq. (4.14).
-D0:5D 0:4
D ref d50s Ss - Sfph X s C v
RH,s = B00 0:08 ð9:18Þ
Dref d50ref 1:65 0:15
X f þ X p þ X h C v ð Ss - Sl Þ
Sfph = Sl þ ð5:23Þ
1 - Cv 1 - X f - X p - X h
1
distribution for a sand 0.9
product with d50 = 0.8 mm 0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0.8
0.1 1 10
Particle size (mm)
9.2 Settling Slurries with Newtonian Liquids 285
B″ again expresses the frictional reduction by the finer particles and is constructed
in a similar way as C″, and Vsm,s is the deposition limit velocity for the ds-sized
particle (ds = 0.015DD).
Sfp = Sl þ 0 = 1
C 00 = 1 - 0 = 1
Sfp = Sl þ 0 = 1
The heterogeneous component is then calculated using Eqs. (5.28) and (9.15):
C 00 = 1 - 0 = 1
ð - 0:51 Þ0:3 0:4
0:3 0:55 2:65 - 1 0:75 0:15
RH,h = 1 0:08 = 0:057
1 1 1:65 0:15
The flow rate of 100 m3/h = 0.028 m3/s, and taking the pump discharge diameter
of DD = 0.103 m gives a mixture velocity Vm = 3.334 m/s. The deposition limit
velocity Vsm,s = 1.4 m/s from Fig. 4.8 or Eq. (4.14). The fully stratified component
can then be calculated using Eq. (9.18), with supporting Eqs. (5.32), (9.16), and
(9.17):
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
V 100,s = 3 1800 9:81 0:103 0:148 = 6:44 m=s
0:5
6:44 - 3:334
B00 = 1 - ð0:5 0:75Þ = 0:706
6:44 - 1:4
9.3 Non-Newtonian Slurries 287
HR = 1 - 0:077 = 0:923
With very fine particles or very high solids concentrations, most slurries are practi-
cally non-settling and, in industrial applications, many behave in a non-Newtonian
way, giving widely varied pump performance effects. In this case, the solids effect is
mainly related to the rheological behavior. As with Newtonian viscous fluids, small
pumps are affected more than large ones when pumping highly viscous,
non-Newtonian slurries, and the derations on efficiency are normally larger than
those on the head. Some typical results are shown in Table 9.1.
The greater part of pumped industrial slurries fall into the intermediate area
between heterogeneous (settling) and homogeneous (non-settling) types of slurries.
A typical example where rheology-related methods apply is mine tailings transport
with average particle sizes of 10–100 μm and maximum sizes of up to a few
millimeters, when pumped at a high volumetric solids concentration of 0.4–0.5
(i.e., thickened tailings). The tailings thickening and pipeline transportation place-
ment chain has been described in Chap. 6, and details about rheology-based pseudo-
homogenous pipeline friction loss design analyses and methods were covered in
Chaps. 3 and 6.
Extensive loop tests for thickened tailings products have been performed at the GIW
Hydraulic Laboratory (Sellgren et al. 2011; Furlan et al. 2013, 2014, 2016a, b). An
Table 9.1 Typical pump solids effects for selected non-settling industrial slurries at flow rates near
QBEP. Impeller size 0.63 m
Product Bauxite Red mud Kaolin Phosphate clay
Typical slurry density ratio: Sm 1.6 1.55 1.35 1.1
Head reduction RH (%) 2 3 3 2
Efficiency reduction Rη (%) 5 6 6 3
288 9 Effect of Solids on Pump Performance
9.3.2 Modeling
Heywood and Slatter (2019) reviewed published pump derating methods for
non-Newtonian fluids and slurries and found that no commonly accepted method
has yet emerged. The semi-empirical methods for non-Newtonian slurries by Walker
and Goulas (1984) and Graham et al. (2009) will be described below. Because of the
uncertainty associated with these methods and the difficulties associated with char-
acterizing non-Newtonian slurries in the first place, test work is recommended to
assist in refining these predictions for cases where the head derations may be high
(e.g., HR < 0.9) and where any uncertainties in the magnitude of the deration will
present significant risks, such as in the design of a long pipeline system with multiple
pumps and high total head.
The 4-Component procedure for settling type slurries may serve as an indicative
or complementary method for highly concentrated tailings slurries when the fine-
particle portion is not very rheologically active (i.e., with no or a small clay content).
Most products in Table 9.2 have Xf in the 0.20 to 0.50 range, Xp of 0.80 to 1.0, and Xh
of 0 to 0.20. With a Cv of 0.4 to 0.45, then the RH, according to the 4CM, will be in
the range of 0.05 to 0.10, similar to the measured values. However, unlike many
settling slurries, Rη will normally be higher for concentrated tailings slurries, which
must be considered when sizing the pump driver.
9.3
Table 9.2 Derating results for thickened tailings slurry loop testsa
D (m) Slurry name Slurry makeup N (rpm) HR ER
Non-Newtonian Slurries
ρm ωD2
Reω = ð9:19Þ
ηB
where D is the impeller diameter, ρm the slurry density, ω the rotary speed (in radians
per second), and ηB the Bingham tangent or plastic viscosity.1 Correlations for HR
and ER based on the Walker and Goulas data were subsequently developed by
Paterson and Cooke, as shown in Fig. 9.12.
1.00
Head Ratio
0.95
Head and Efficiency Ratio (HR, ER)
0.90
0.85
0.80
0.75
Fig. 9.12 Head and efficiency derate correlations (Paterson & Cooke derate chart developed from
the work of Walker and Goulas)
1
Note that ηΒ is not an actual viscosity but only a coefficient of the Bingham model, which only
approximates the actual viscosity at high shear rates, where it is asymptotic.
9.3 Non-Newtonian Slurries 291
Values of this Reynolds number less than about 106 give significant reductions in
efficiency, which will generally be larger than the corresponding head derate.
Results for pumps with impeller diameters from 0.4 to 1 m by Sery and Slatter
(2002) and Xu et al. (2002), with Bingham-type slurries with yield values of
100–200 Pa, support the method. An analysis of the data in Table 9.2 shows an
accuracy of the head deration to within ±5 percentage points and a general over-
prediction of the efficiency deration in the range of 5–15 percentage points. Since the
head deration is most important where pump-system interaction is concerned, and a
safety factor on efficiency mainly affects the installed driver power, the method
remains useful and still finds widespread application.
The pump rotary speed dependence on both head and efficiency for the results in
Table 9.2 was generally larger than predicted and increased with the amount of very
fine particles.
It should be noted that the modified pump Reynolds number does not consider the
yield stress and relies solely on the tangent viscosity for rheological characterization
of the slurry. Since the viscosity of the slurry approaches the plastic viscosity at low
yield stress and/or high strain rates, it may be inferred that the model is best suited for
lower yield stress slurries (i.e., < 100 Pa), higher flow rates, and/or larger pumps
running at higher speeds.
As shown in Fig. 9.2, increasing the fluid’s viscosity derates the pump’s perfor-
mance. For Newtonian fluids, where the viscosity is insensitive to shear rate, the
calculation of the derate is well established by methods such as the Hydraulic
Institute guideline for effects of liquid viscosity on performance (ANSI/HI 9.6.7
(2021)). Such a direct approach is not possible for non-Newtonian fluids and slurries,
as their viscosity varies with shear rate and consequently varies with location within
the pump, impeller speed, and flow rate. This complexity means that very few
studies have been made on pump performance with non-Newtonian fluids, and it
is difficult to draw generalizations from those that do exist, due to the wide
variability of non-Newtonian fluids. The analysis could be simplified if one could
identify some characteristic fluid viscosity within the pump that would allow the
application of a Newtonian pump performance derate model, even if this viscosity
varied according to the pump flow rate.
For laminar pipe flows, it is reasonable to assume that the flow within a normally
sized pump will also be laminar, since the flow areas within the pump are often of a
similar size, or smaller, than the associated pipeline. Graham et al. (2009) proposed
that the flow through the complex geometry of a spinning impeller could be
approximated to laminar flow through a simple duct of hydraulic diameter (Dh),
which was itself unknown, but might be obtained from a large number of carefully
controlled pump tests for a variety of true non-Newtonian fluids. The bulk velocity
through such a duct is then simply:
292 9 Effect of Solids on Pump Performance
4Q
V= ð9:20Þ
πD2h
The hydraulic diameter was then defined based on the area (A) and perimeter (P) of a
typical impeller outlet:
4A 4wπD
Dh = = ð9:21Þ
P 2ðπD þ wÞ
D
w≈ ð9:22Þ
4
This value should be treated with care until more pump types can be tested. Using
this correlation, the authors were able to predict the deration of several pump/slurry
combinations. Note that the method is only suitable for laminar flow. For turbulent
pipe flows, the authors suggested employing the Walker and Goulas method, after
replacing the Bingham plastic viscosity used in Eq. (9.19) with a viscosity obtained
at a suitably high shear rate, to represent the wall shear rates in turbulent flow. By
using this approach, the Walker and Goulas method is no longer restricted to
Bingham plastic flows, and the Reynolds number is now a true ratio of inertial to
viscous forces. From their tests, they established that wall shear rates_γ 0 ≥ 4000s - 1
were sufficiently high to represent turbulent flow.
The Graham and Pullum method also predicts a pronounced drop in performance
at very low flow rates, as is sometimes seen in the experimental data. An example is
given in Fig. 9.13, where the method is used to predict such behavior as measured by
Walker and Goulas. In the model, this “droop” in the head curve is the result of an
increase in the characteristic viscosity with decreasing shear rate. However, Furlan
et al. (2013, 2014) have proposed that the corresponding effect in the experimental
results, which is present in some laboratory data and absent from others, is due to “air
lock” (i.e., a blockage in the suction of the pump caused by air entrainment), since
2
Note the authors used a method cited in an earlier (1969) version of the Hydraulic Institute
standard which also included small pump data, which might be considered equivalent to flow
close to the shutoff condition in a normal-sized slurry pump.
9.3 Non-Newtonian Slurries 293
the removal of entrained air in closed loop laboratory systems with highly viscous
slurries is a common problem. Air lock tends to become more severe at both low
flow and high viscosity, since both of these conditions enhance the ability of the air
to collect and remain in the pump suction, and can lead to a similar drooping curve.
In a series of tests handling both clay-based and non-clay tailings slurries with
yield stresses from 100 to 500 Pa and beyond, Furlan et al. passed the slurries over a
sharp-edged weir to break any bubbles, thus releasing the air before it re-entered the
pump (See Fig. 11.21). Subsequently, no drooping of the head curves was observed.
However, Graham et al. (2009) exercised great care to ensure that the fluids tested in
their study were accurately characterized with regard to rheology, and that the tests
were conducted in such a manner that air was not entrained or retained within the
fluids being tested. Ultimately, further research is needed to fully establish the
mechanisms involved in generating these drooping head curves. Fortunately, the
question is somewhat academic, since the low values of Q/QBEP at which it occurs
are generally avoided outside of research laboratories, and most industrial installa-
tions in service today are fitted with variable-speed drivers capable of overcoming
the ill effects of high head losses during start-up, when flows are low. Still, industrial
pump feeding practice is less controlled than in research laboratories, meaning that
air entrainment may be present, so whether by viscosity effect or air-lock, these
droops in head performance at low flow rate may be experienced and should be
allowed for when selecting the pumping capacity.
294 9 Effect of Solids on Pump Performance
Section 8.3 dealt with cavitation and related problems with pump suction perfor-
mance when pumping water. The NPSHR specified by the pump manufacturer is
almost always based on tests using clear water. This can be considered equal to the
NPSHR on slurry, if expressed in head of slurry, and assuming that the slurry
approximates an equivalent liquid behavior (i.e., the solids effect on head is small
or zero). When significant solids effect on head exists, some derate in the NPSHR
performance may also be expected, although few data exist on this topic. Standard-
ized closed vacuum tank results by Herbisch (1975) for sands at various slurry
densities support an equivalent liquid treatment. Roudnev (2004) reported similar
sand and ore results and discussed correlations for Bingham-like fiber and ore
suspensions. On the other hand, in an attempt to simulate field suction-side condi-
tions in a phosphate dredging pit, Whitlock et al. (2001) induced cavitation with an
open-tank, valved suction, laboratory arrangement with a 1.4 m diameter impeller
pump. The NPSHR in meters of slurry was about 50% higher than the corresponding
specified water value when pumping various sands for Cv values of up to 24%.
However, being open tank tests, it is possible that entrained air or the suction side
valving affected the measurements. In general, it is recommended that some correc-
tion (increase) to NPSHR be made proportional to the head reduction factor. Using
the head reduction factor directly as an NPSHR increase factor would appear to be an
overly conservative, although safe approach. Using one-half of the head reduction
factor is likely closer to the actual value.
For viscous Newtonian liquids, ANSI/HI 9.6.7 suggests that NPSHA should
be > (NPSHR/HR), in addition to the margin normally used for water and this
same adjustment would be applied to a non-Newtonian slurry when using the
Graham and Pullum method. Furlan et al. (2014) used a novel sealed-tank arrange-
ment to investigate the suction performance of a concentrated tailings slurry at
volumetric concentrations above 0.5 with a 200–300 Pa yield stress, (slurries A
and B in Table 9.2). The measured NPSHR values were within 20% of the clear
water results.
This case study is a continuation of Case Study 5.1, where the pipeline characteris-
tics were calculated for a horizontal pipeline of 350 m in length. The pipeline inner
diameters, operating velocities, and pumping head requirements were determined,
together with the specific energy consumption (SEC) to overcome the pipeline
friction losses for 500 tons/h, as summarized in Table 9.3.
9.5 Case Studies 295
Table 9.3 Calculated pipeline operating data from, Table 5.2, Case Study 5.1
D(mm)/Cv Vm,500 (m/s) im (-) SEC (kWh/ton-km) Hm,frict (m slurry)
202.7/0.35 4.64 0.172 0.505 38.1
254.5/0.25 4.12 0.122 0.502 30.2
304.8/0.15 4.79 0.091 0.622 25.4
Fig. 9.14 Performance curves with duty points for pumps from Table 9.4
The objective here is to select suitable pumps for the three pipelines, establish
their operating points (including pump speed and efficiency), and then to calculate
the electric power required to deliver 500 tons per hour of dry solids over the
horizontal distance of 350 m. The selection of a suitable pump for each of the
three pipelines, together with the evaluation of the total energy requirements, form
the basis for a comparison of the three systems.
For the pump selection, flow rates in m3/h are calculated from the pipeline
diameter and corresponding velocity data of 4.64, 4.12, and 4.79 m/s from
Table 9.3, giving flow rates of 539, 754, and 1258 m3/h, respectively. These values,
together with the tabulated required pumping heads (Hm) in meters of slurry, form
the primary input for pump selection. In practice, these flow rates and heads are
compared to the available slurry pumps. This is done by comparing the manufac-
turer’s multi-speed pump performance curves to the required operating conditions,
either manually or, more commonly, using a computer program. In both cases,
pumps operating near 0.8 QBEP at the desired flow and head are often considered
the best selections, since this relative flow rate often optimizes the parameters of
pump efficiency, NPSHR, and wear performance, while allowing room for future
increases in flow rate and production. The actual practice of pump selection will be
more complex, as covered in subsequent chapters, but for the purpose of the present
example, we will assume the 0.8 QBEP criteria. This leads to the selection of three
different pumps with impeller diameters of 0.61, 0.66, and 0.71 m, corresponding to
the concentrations by volume of 0.35, 0.25, and 0.15, respectively. Their perfor-
mance curves and operating points are shown in Fig. 9.14.
Using the 4-Component pump solids effect model described above, the head
reduction factors are calculated for each of the three pumps at their individual
296 9 Effect of Solids on Pump Performance
impeller and discharge diameters (D and DD) and operating conditions of solids
concentration and particle size distribution. All particles are larger than 0.2 mm;
therefore, Xf and Xp = 0. The pumps all have discharge diameters greater than
150 mm; therefore, the particle size boundaries for the Xs fractions (0.015DD) are all
greater than 2.25 mm, which means that Xs = 0 and Xh = 1 because the maximum
particle size is about 1 mm. The d50 = 0.47 mm, and since only Xh solids are present,
this is also = d50h. These conditions are very similar to those of Example 9.2, part
1, with a heterogeneous contribution only. Therefore, as in that previous example:
C 00 = 1 - 0 = 1
ð - 0:51Þ0:61 0:4
0:61 0:47 2:65 - 1 1 0:35
RH,h = 1 0:08 = 0:160
1 1 1:65 0:15
ð - 0:51Þ0:66 0:4
0:66 0:47 2:65 - 1 1 0:25
RH,h = 1 0:08 = 0:113
1 1 1:65 0:15
ð - 0:51Þ0:71 0:4
0:71 0:47 2:65 - 1 1 0:15
RH,h = 1 0:08 = 0:067
1 1 1:65 0:15
The pump speed and efficiency are calculated according to the procedure example
in Sect. 9.1 in association with Fig. 9.5. It is assumed that HR = ER. The slurry
References 297
densities for the three concentrations 0.35, 0.25, and 0.15 are calculated with 1000 .
(1 + Cv . (2.65 - 1) to give 1578, 1413, and 1248 kg/m3, respectively.
The pressure requirements are calculated as the slurry density in kg/m3 times the
pumping heads in meters of slurry times g. Thereafter, the power (W) is calculated
with the pressure times flow rate (m3/s) divided by the total pump efficiency,
including an assumed drive efficiency of 95%. Finally, the total system SEC is
calculated with the power (kW) divided by the tonnage per hour (ton/h) and the
pipeline length of 0.35 km. A summary of the resulting total system evaluation is
given in Table 9.4:
References
T. Engin, M. Gur, Comparative evaluation of some existing correlations to predict head degradation
of centrifugal slurry pumps. ASME J. Fluids Eng. 125, 149–157 (2003)
L.C. Fairbanks, Effects on the characteristics of centrifugal pumps. Trans. Am. Soc. Civ. Eng. 107,
1564–1575 (1941)
J.M. Furlan, R.J. Visintainer, A. Sellgren, Pumping a 100 – 600 Pa yield stress clay slurry with a
centrifugal pump. Presented at the 16th International Seminar on Paste and Thickened Tailings,
Belo Horizonte, 2013
J.M. Furlan, R.J. Visintainer, A. Sellgren, Centrifugal pump performance on a high-yield stress
tailings slurry. Presented at the 17th International Seminar on Paste and Thickened Tailings,
Vancouver, 2014
J.M. Furlan, R.J. Visintainer, A. Sellgren, Centrifugal pump performance when handling highly
non-Newtonian clays and tailings slurries. Can. J. Chem. Eng. 94(6), 1108–1115 (2016a)
J.M. Furlan, R.J. Visintainer, A. Sellgren, Centrifugal pump performance when handling
non-Newtonian slurries with different particle sizes and densities. Presented at the 18th Inter-
national Seminar on Paste and Thickened Tailings, Santiago, 2016b
R.G. Gillies, K.B. Hill, M.J. McKibben, C.A. Shook, Solids transport by laminar Newtonian flows.
Powder Tech. 104, 69–277 (1999)
L.J.W. Graham, L. Pullum, L. Slatter, G. Sery, M. Rudman, Centrifugal pump performance
calculation for homogeneous suspensions. Can. J. Chem. Eng. 87, 526–533 (2009)
A. Haider, O. Levenspiel, Drag coefficient and terminal velocity of spherical and non-spherical
particles. Powder Tech. 58, 63–70 (1989)
J.B. Herbisch, Coastal and Deep Ocean Dredging (Gulf Publishing, Houston, 1975)
N. Heywood, P. Slatter, Review of methods for the deration of centrifugal pumps for
non-Newtonian slurries. Presented at the 19th International Conference on Transport and
Sedimentation of Solid Particles, Cape Town, RSA, 2019
K.A. Kazim, B. Maiti, P. Chand, A correlation to predict the performance characteristics of
centrifugal pumps handling slurries. Proc. Inst. Civ. Eng. A 211, 147–157 (1997)
L. Martinson, R. Martinson, R. Cooke, Pipe loop tests at Codelco pilot plant. Presented at the 16th
International Seminar on Paste and Thickened Tailings, Belo Horizonte, 2013
R.E. McElvain, High pressure pumping. Skill. Min. Rev. 63(4), 1–14 (1974)
W. Mez, The influence of solids concentration, solids density, and grain size distribution on the on
the working behavior of centrifugal pumps. Presented at the 9th International Conference on the
Hydraulic Transport of Solids in Pipes, Rome, 1984
D.M. Newitt, J.F. Richardson, M. Abbot, R.B. Turtle, Hydraulic conveying of solids in horizontal
pipes. Trans. Inst. Chem. Eng. 33, 93–113 (1955)
F. Ni, W.J. Vlasblom, A. Zwartbol, Effect of high solids concentration on characteristics of a slurry
pump. Presented at the 14th International Conference on Slurry Handling and Pipeline Trans-
port, Maastricht, the Netherlands, 1999
M.C. Roco, M. Marsh, G.R. Addie, M. Maffett, Dredge pump performance prediction. J. Pipelines
5, 171–190 (1986)
A.S. Roudnev, Slurry pump suction performance considerations. Presented at the 16th International
Conference on the Hydraulic Transport of Solids, Santiago, Chile, 2004
A. Sellgren, Performance of centrifugal pump when pumping ores and industrial minerals.
Presented at the 6th International Conference on the Hydraulic Transport of Solids in Pipes,
Canterbury, 1979
A. Sellgren, G.R. Addie, Effect of solids on large centrifugal pump head and efficiency. Presented at
the Central European Dredging Association Dredging Days, Amsterdam, 1989
A. Sellgren, L. Vappling, Effects of highly concentrated slurries on the performance of centrifugal
pumps. Proceeding of the ASME International Symposium on Slurry Flows FED 38:143–148,
(1986)
A. Sellgren, D. Luther, G. Addie, Effect of coarse and heavy solid particles on centrifugal pumps
head and efficiency. Presented at the Canadian Chemical Engineering Conference,
Edmonton, 1997
References 299
A. Sellgren, G. Addie, L. Whitlock, Using centrifugal pumps for highly concentrated tailings
slurries. Presented at the 15th International Conference on the Hydraulic Transport of Solids,
Banff, 2002
A. Sellgren, A. Mustafa, GA Addie, L. Whitlock, Performance of different centrifugal pump
impeller configurations when pumping thickened tailings. Presented at the 14th International
Seminar on Paste and Thickened Tailings, Perth, Australia, 2011
A. Sellgren, R. Visintainer, J. Furlan, Centrifugal slurry pump performance deratings-a coherent
approach. Presented at the 20th International Conference on Hydrotransport, Melbourne, 2017
G. Sery, P.T. Slatter, Pump derating for non-Newtonian slurries. Presented at the 15th International
Conference on the Hydraulic Transport of Solids, Banff, 2002
C.A. Shook, R.G. Gillies, M. McKibben, Derating of a centrifugal pump by large solid particles.
Proceeding of the 8th International Freight Pipeline Society Symposium, University of Pitts-
burg, Pittsburg, pp. 144–151 (1995)
A.J. Stepanoff, Pumps and Blowers: Two-Phase Flow (Wiley, New York, 1965)
P.M. Vercruijsse, F. Corveleyn, The solids effect on pump and pipeline characteristics- Keeping up
with present trends in the dredging industry. Presented at the 15th International Conference on
the Hydraulic Transport of Solids, Banff, 2002
R. Visintainer, A. Sellgren, J. Furlan, G. McCall, Centrifugal pump performance deratings for a
broadly graded (4-Component) slurry. Presented at the 18th International Conference on
Transport and Sedimentation of Solid Particles, Prague, 2017
R. Visintainer, A. Sellgren, V. Matousek, G. McCall, Testing and modelling of diverse iron ore
slurries for pipeline friction and pump head derate. Presented at the AusIMM Iron Ore
Conference, Perth, 2021
R. Visintainer, G. McCall, A. Sellgren, V. Matousek, Large scale, 4-component, settling slurry test
for validation of pipeline friction loss and pump head derate models. WEDA J Dredging,
Western Dredging Association, Temecula, 2022
J.J. Vocadlo, J.K. Koo, A. Prang, Performance of centrifugal pumps in slurry service. Presented at
the 3rd International Conference on the Hydraulic Transport of Solids in Pipes, Golden, 1974
C.I. Walker, A. Goulas, Performance characteristics of centrifugal pumps when handling
non-Newtonian homogeneous slurries. Proc. Inst. Mech. Eng. 198A, 41–49 (1984)
L. Whitlock, A. Sellgren, K.C. Wilson, Net positive suction head requirement for centrifugal slurry
pumps, in Handbook of Conveying and Handling of Particulate Solids, ed. by A. Levy,
H. Kalman, (Elsevier, 2001), pp. 491–497
K.C. Wilson, G.R. Addie, A. Sellgren, R. Clift, Slurry Transport Using Centrifugal Pumps, 3rd
edn. (Springer, New York, 2006)
D. Wolfe, The effect of large particles on oil sand hydrotransport slurry pump performance.
Presented at the 19th International Conference on Hydrotransport, Golden, 2014
J. Xu, R. Tipman, R. Gillies, C. Shook, Centrifugal pump performance with Newtonian and
non-Newtonian slurries. Presented at the 15th International Conference on the Hydraulic
Transport of Solids, Banff, 2002
Chapter 10
System Stability and Operability
10.1 Introduction
Earlier chapters have referenced the need to match the head characteristics of the
pipeline with those of the pumps to ensure stable operation of a slurry system. This
chapter illustrates how we can apply an understanding of slurry behavior to an
analysis of system operability by examining both steady-state operation and transient
responses to changes in slurry properties. The focus is on two common sources of
variation: changes in solids concentration with constant particle size distribution,
and changes in particle size at fixed solids concentration. However, the reasoning
applied can be extended to more complex effects. Settling slurries are considered in
Sects. 10.2 to 10.4 while non-Newtonian slurries are discussed in Sect. 10.5. The
application of specific energy consumption (SEC) concepts to system design is
treated in Sect. 10.6.
For the operator of a slurry system, it is important to understand how the system
will respond to variations in the properties of the slurry fed to it. It is also important
for the designer to appreciate fully the implications of variations in particle size,
concentration, and throughput. As for all handling and processing operations, the
design must reflect the range of variability that is anticipated. For slurry systems, this
implies finding the economic balance between strict control and optimization of
slurry consistency on the one hand, and, on the other hand, allowances for slurry
variability and non-optimal conditions. For example, with long-distance transport
pipelines, the dominant capital and operating costs are usually associated with
conveying. Therefore, much effort is devoted to preparing and controlling slurry
properties. At the other extreme, short in-plant conveyors and dredging systems may
need to accommodate wide and sometimes uncontrollable variations in solids size,
rheology, and throughput. In actual practice, significant savings may be possible by
limiting operational variability and improving stability, particularly in energy effi-
ciency; however, the ability to handle whatever variability will arise must be
considered.
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 301
R. Visintainer et al., Slurry Transport Using Centrifugal Pumps,
https://doi.org/10.1007/978-3-031-25440-6_10
302 10 System Stability and Operability
Vm (a)
high conc.
jm medium conc.
low conc.
water
Vm (b)
own density. Absolute values are deliberately not given on the coordinate axes, so
the system curves may be interpreted as total head loss for a fixed pipe length. Flow
rate (velocity) is always driven to the intersection of the pump curve and system
curve, and the stability of this intersection point is a key factor in determining overall
system stability.
The first and most obvious result of the equivalent liquid model, seen in
Fig. 10.1b, is that all of the pump and pipeline characteristic curves collapse to
single lines when each is expressed as head of slurry at its own density. This means
the system will operate at a fixed velocity, regardless of changes to the concentration
(i.e., mixture density) of the slurry. This assumes, of course, that the pump speed
remains constant. It also assumes that the concentration is constant at all points
throughout the pump and piping (i.e., any change in concentration occurs slowly
enough that the differences throughout the system are negligible).
To examine the effects of transient operation, where slurry properties may vary
between pump and system, we must examine the im curves in Fig. 10.1a. The jm
curves assume a common slurry density throughout pump and piping, but im head,
expressed in height of standard water, will be proportional to the pressure gradient
within the system and will allow us to consider a variable slurry density between
pump and system. If we examine the intersections in the im curves, we can see how
fluctuations in concentration between pump and system may result in transient flow
rate fluctuations, the flow falling to a lower value when pump concentration is low
and system concentration high (point 1), or to a higher value when the opposite is
true (point 2).
In the equivalent liquid case, system stability is usually maintained, even with
such fluctuations, since the system curves are monotonically increasing, and the
centrifugal pump curves are (normally) monotonically decreasing. The flow rate
does fluctuate, but it remains within defined limits, as illustrated by the shaded area
in Fig. 10.1. Since the slurry solids do not settle at most velocities, the system
operates without interruption.
Other potential variations include changes in the system static head (perhaps due
to changes in sump height), which will raise or lower the curves; and viscosity
variations, which will change the slopes of both the pump and system curves.
However, a similar analysis can be made, and in the case of the equivalent liquid,
stability is usually maintained. Equivalent liquid behavior also provides stability
with respect to the power consumption, which, all other things being equal, will be
directly proportional to the slurry density.
Figure 10.2 shows typical system characteristics for a heterogeneous settling slurry
at three delivered concentrations. As before, two plots of the same data are given,
where the friction gradient is expressed either as head of standard water (im) or head
304 10 System Stability and Operability
im
c
b
a
d
e
Vm
high conc.
jm medium conc.
low conc.
water
Vm
Fig. 10.2 System and pump characteristics for the heterogeneous flow of a settling slurry at various
delivered concentrations, given in terms of head of standard water (im) and head of slurry ( jm)
of slurry ( jm), and only the frictional contribution for horizontal transport is
considered.
The form of these system characteristics is typical of the heterogeneous flow of a
settling slurry, as discussed in Chap. 5, and the curves all show minima typical of
that flow regime. As implied by the relations given in Chap. 5, the position of the
minimum moves to higher velocities as the solids concentration increases. Also, as is
typical for heterogeneous slurries, this minimum friction gradient occurs at velocities
above the deposition limit velocity. Note that the curves no longer collapse when
expressed as jm, since the solids effects in both the pump and the pipeline are no
longer directly proportional to the slurry density.
Assuming the centrifugal pump operates at a constant speed, the total developed
head measured in terms of delivered slurry density ( jm) decreases slightly with
increasing concentration, due to the effect of the solids on pump performance
discussed in Chap. 9. However, the effect of density is usually greater than the
solids effect, so that the pump head, expressed as im (i.e., the water column
equivalent to the discharge pressure of the pump), increases with slurry concentra-
tion. The increase is less than in direct proportion to Smd by the amount of the
aforementioned pump solids effect.
10.3 Effect of Solids Concentration in Settling Slurries 305
The transient behavior is of central interest. Consider the case where the system
has been operating steadily at the medium concentration point “a,” and the slurry
presented to the pump suddenly changes to a higher concentration. Referring to the
upper graph, the system characteristic remains at the medium concentration, but the
pump is now handling a higher-density slurry, so that its head curve increases to the
higher concentration characteristic. The immediate effect is to shift the system
operating condition to point “b,” increasing both the mean slurry velocity and the
power drawn by the pump. As the higher solids concentration propagates along the
line, the system resistance moves up to the higher characteristic, so that the velocity
decreases and system operation moves up to point “c.” Note that the velocity has
returned to a value near to the original condition, although slightly lower. Con-
versely, if the system has been operating steadily at point “a” and the slurry entering
the pump is suddenly reduced to a low concentration, the mixture velocity is reduced
as the system moves to point “d” and on again to point “e” as the lower concentration
moves through the system. Again, the flow rate has returned to a value near to the
original condition, although it is now slightly higher.
At first glance, this case appears similar to the equivalent liquid case, with a
similar range of potential operating conditions indicated by the shaded area. Indeed,
if we examine the jm curves, we can see that the system is operating in an area where
the system characteristic curves converge. For heterogeneous settling slurries, this
convergence often occurs near the area where the slurry system curves cross the
characteristic curve for the conveying fluid (in this case, water). In this region, jm is
approximately equal to jf, which is the hallmark of equivalent liquid behavior, so this
is sometimes called the equivalent liquid point, although it is not always a well-
defined, single point. Operating in this region provides some automatic compensa-
tion similar to that provided in the equivalent liquid case.
While operation near the equivalent liquid point may be optimal for system
stability, it is not always optimal for energy efficiency, as discussed in Sect. 10.6.
For a heterogeneous slurry, the energy efficiency is often optimized by operation
closer to the minimum friction velocity, so in Fig. 10.3, let us consider the same
piping system and slurry as in Fig. 10.2, but with a pump speed selected to operate
closer to the minimum friction velocity. In this region, the flow transitions from pure
heterogeneous to a mixture of stratified and heterogeneous flow, as described in
Chap. 3, leading to the typical upward-sloping curve at low velocity and the creation
of the minimum friction point.
Examining the jm curves, we see that a broader range of intersections leading to a
wider range of pipeline velocities is now possible. In fact, there appears to be no
intersections at all between the pump and system curves at the highest concentration.
Examining the im curves, consider the effect of increasing concentration from the
medium to the high characteristic. As before, the effect on the pump occurs before
the new concentration has propagated along the pipeline, so assuming a start at point
“a,” the immediate effect is to shift operation to point “b.” However, as the higher
concentration propagates through the system, friction losses increase, and the system
resistance curve moves up towards the high-concentration characteristic. As the
characteristic rises, the intersections point falls to an ever-decreasing velocity.
306 10 System Stability and Operability
im
plug
b
c a
water
Vm
high conc.
jm medium conc.
low conc.
water
Vm
Fig. 10.3 System and pump characteristics of Fig. 10.2, but with a different pump speed
Eventually, the system resistance reaches the dashed curve and the pump can no
longer generate sufficient head to maintain flow at any velocity. Velocity falls off
abruptly into the deposit region, and with more solids constantly entering the system
and settling out, the line becomes “plugged.” At this point, reducing the solids
concentration entering the system, even to the point of pumping fluid alone, cannot
clear the plug; higher pump speeds are needed. Alternatively, a high-concentration
slurry of fine particles may shift the deposit by providing increased pressure from,
and reduced solids effect losses on, the pump. If a variable-speed pump or a high-
concentration, fine-particle slurry is not available, the only recourse will be to open
up the line at some intermediate point and flush the solids out.
A similar condition of plugging may occur if the concentration of the slurry
entering the pump is suddenly reduced to the low-concentration level. The pump
characteristic falls to the low-concentration line, but the system characteristic is still
at a medium concentration and, once again, no intersection is available as operation
attempts to move to point “c.” Once again, the velocity will fall off and deposition
will begin. In this case, however, recovery is possible if the concentration entering
the pump is again increased before the pipeline has the opportunity to develop a
significant deposit, or otherwise clog at an elbow or other obstruction. Alternatively,
the reduction in velocity may result in enough solids falling out of suspension that
10.3 Effect of Solids Concentration in Settling Slurries 307
the system curve drops to a lower concentration value, once again establishing an
equilibrium, albeit with a bed of solids now in the pipeline, a condition that can lead
to other problems.
Some general conclusions can be drawn from the foregoing discussion. Compar-
ing the system and pump characteristics is essential because it enables a qualitative
but very informative assessment of operating stability. Operating velocities should
be high enough to allow for expected transients in concentration, static head, pump
solids effects, etc., without falling below the system minimum friction velocity. A
variable-speed pump driver may provide critical flexibility to handle transients, but
there will still be limits that must be considered.
Figure 10.3 also illustrates why the velocity at the limit of stationary deposition is
often unimportant for heterogeneous settling slurries: While operation led to a
“plugged line,” the cause was poor matching of the pump and system characteristics,
rather than operation too close to the deposition limit velocity. The minimum friction
im
d
a
c e
b
Vm
coarse
jm medium
fine
water
Vm
Fig. 10.4 Settling slurry system and pump characteristics at constant concentration with variable
particle size
308 10 System Stability and Operability
velocity (which for many heterogeneous slurries is above the deposition limit
velocity) was the limiting factor. This explains why field data (and some laboratory
data) indicate deposition velocities much above the values estimated from the
analysis in Chap. 4: They actually correspond to the limit of stable operation with
centrifugal pumps, rather than the limit of operation without a stationary deposit. In
practice, centrifugal pumps permit operation near the deposition point only for
relatively fine particles.
Figure 10.4 shows system and pump characteristics for the case where the slurry
concentration remains constant, but three different particle sizes are considered
designated “fine,” “medium” and “coarse.” Note that the effect of increased particle
size is similar (but not identical) to the effect of increased concentration. The curves
for the medium-particle slurry are identical to those for the medium concentration in
the earlier figures, and the conversion between im and jm has been carried out using
the same medium value of Sm.
The coarse slurry in this case is tending towards the fully stratified flow regime.
Because there is more stratification, its corresponding system characteristic shows a
pronounced minimum, and at a velocity higher than seen for the medium slurry.
Similarly, the crossover velocity is also higher for the coarse slurry, and there is no
longer a clear equivalent liquid point between the curves. The fine slurry in this case
shows no minimum friction point and approaches the case of pseudo-homogeneous
flow and equivalent liquid behavior as described above.
Since concentration is not changing, the spread between pump curves on the im
plot does not change much relative to the jm plot, with only a small loss in pump head
in both cases due to the increasing particle size. In both cases, an increase in the
particle size results in a reduction in pump performance and, therefore, a drop in
pump head. Assuming a starting condition of medium particle size throughout,
introducing coarse particles into the pump decreases pump head, thereby reducing
velocity and moving the operating condition from point “a” to point “b.” As the
coarse solids move through the system, the friction losses increase, and the velocity
reduces further to point “c.” In this case, the system continues to operate. If,
however, we were attempting to operate closer to the minimum friction velocity,
as illustrated above for the case of variable concentration, a similar plugging issue
could develop, as one can imagine by comparing the curves.
Best practice in settling slurry system design is to always check both the mini-
mum friction and stationary deposition velocities and plan to operate, with some
margin, above whichever is greater.
The cases presented here to illustrate the analysis of system operability have been
based on the relatively flat head characteristics typical of centrifugal slurry pumps
with relatively low specific speed. Higher specific speed designs will give a more
steeply falling pump characteristic, which may theoretically improve operating
10.5 Non-Newtonian Slurries 309
Figure 6.2 in Sect. 6.2 shows, in simple generalized terms, the effect of changes in
solids concentration on the resistance of a piping system to flow of a homogeneous,
non-Newtonian slurry. This kind of behavior is again shown schematically in
Fig. 10.5, with the characteristic curves expressed as head of slurry ( jm). As with
the previous examples, only the frictional contribution is shown (i.e., the curves refer
to horizontal transport). At velocities sufficiently high for the slurry to achieve
turbulent flow, the head lost in flowing through the system can be estimated by
treating the slurry as a fluid with an effectively constant friction factor. This leads to
the “turbulent” system curves in Fig. 10.5, for which the influence of the solids
concentration is small, since turbulent flow is largely inertial and so mainly responds
to the variations in density, rather than to viscosity.
high conc.
jm medium conc.
low conc.
water
pump 2
pump 1 e
c
b
a
Vm
Fig. 10.5 System and pump characteristics for flow of a homogeneous, non-Newtonian slurry at
three concentrations, expressed in head of slurry ( jm)
310 10 System Stability and Operability
However, in laminar flow (which develops at lower velocities in Fig. 10.5), the
system head varies strongly with concentration, since it is dominated by viscous
variation. The transition between laminar and turbulent flow is indicated by the
intersection of the laminar and turbulent portions of the system characteristic.
Therefore, as illustrated in the figure, the transition velocity increases somewhat
with solids concentration.
To examine the operability of such a system with centrifugal pumps, we now
superimpose some pump head characteristics onto the system characteristics. As
shown in Sect. 9.3, the pump head for non-Newtonian slurries often sees only a small
solids effect, provided the yield stress of the slurry is not too high (i.e., below 75 Pa),
so in the present example, the solids effect is neglected and the pump characteristic at
constant speed, expressed in head of slurry, appears as a single curve.
Chapter 6 indicated that the most economic system operation for a homogeneous
slurry may lie close to the transition velocity. Consider first a system that has been
designed for such a case, corresponding to point “a” in Fig. 10.5 with the medium
concentration. The curve labeled “pump 1” represents the characteristic of a pump
selected for this operating point. If concentration decreases to the low value, then
system operation shifts to point “b,” the new intersection of the system and pump
characteristics. Similarly, an increase to high concentration shifts operation to point
“c.” These three conditions span a large variation in velocity. It was shown in
Chap. 6 that relatively small variations in concentration can have strong effects on
slurry rheology, and hence on the laminar flow curves with non-Newtonian slurries.
Thus, when operating in the laminar portion of the system characteristic, the
variations in mean velocity (and hence the solids throughput) are amplified by the
way the system and pump characteristics interact, and steady operation in laminar
flow with a constant pump speed and no flow control valve is only possible if the
slurry consistency is tightly controlled.
For variable solids concentration, two practical alternatives are available. One is
to operate in turbulent flow, such as at point “d” in Fig. 10.5. The curve labeled
“pump 2” shows the head-discharge characteristics of a centrifugal pump selected
for this duty. By operating in the turbulent region of the system characteristic,
variations in slurry consistency can be accommodated without significant changes
in mixture velocity. This is probably the main reason that many designers prefer to
use turbulent flow even for homogeneous non-Newtonian slurries containing no
coarse solids that may settle. An additional reason for operating in turbulent flow
may occur when handling complex slurries (i.e., mixtures of non-Newtonian slurries
with coarse solids), as the turbulence prevents the formation of a stationary bed and
possible associated blockages, as is discussed in Chap. 6. In this case, the design
point is then selected to be in the turbulent range for the highest solids concentration
expected. However, it should be noted that this will not correspond to the lowest
specific energy consumption (SEC), as described in more detail below. This is an
example of the general point that design for variable operation is often incompatible
with design for minimum energy consumption.
The second alternative is variable-speed pump operation, as illustrated by points
“e” and “f” in Fig. 10.5. If the solids concentration increases from medium to high,
10.5 Non-Newtonian Slurries 311
pump 1 e
c
a b
d
Vm
then the pump speed is increased to maintain a constant flow velocity and solids
throughput. The corresponding increase in pump speed raises the pump head
characteristic to the dashed curve shown passing through point “e.” Similarly, if
the concentration falls from medium to low, the pump speed is decreased to give the
characteristic passing through point “f.”
The effects of particle size distribution can be investigated in a similar way.
Generally, decreasing the particle size of a non-Newtonian homogeneous slurry
tends to increase the viscosity. Thus, it is broadly equivalent, in terms of the
coordinates of Fig. 10.5, to increasing the concentration of the system characteristic.
Other variables that affect the slurry rheology (e.g., pH) will produce similar results.
A similar conclusion then follows: If the system will have to handle variations in
slurry rheology, then it must be designed either for turbulent flow or with variable-
speed pumps. If the former option is used, the system must be designed for turbulent
flow at the most “viscous” slurry anticipated.
The preceding discussion concerns long-term system stability in response to slow
changes in slurry properties. To examine the short-term effects of rapid transients, it
is necessary to return to the coordinates in which the system and pump heads are
measured in terms of the density of the carrier liquid. Figure 10.6 shows the pump
and system characteristics of Fig. 10.5 expressed in head of standard water (im). As
with the previous examples, the head (im) is equal to the product of Sm and the slurry
head ( jm). The pump and system curves representing the various densities therefore
expand, and the range of possible pump-system intersections for cases where the
pump and system slurry concentrations differ can again be seen as the shaded areas.
Note that with “pump 2,” the area of intersection remains small, similar to the
equivalent liquid case, or the heterogeneous slurry case when operating near the
312 10 System Stability and Operability
equivalent liquid point. However, with “pump 1,” which interacts with the laminar
region of the system curves, the situation is quite different.
Consider operation at point “a,” with the medium concentration in both pump and
system. If a high concentration is now introduced into the pump, the pump charac-
teristic rises and the velocity increases to the new intersection at point “b.” However,
as the higher concentration propagates through the system, the system characteristic
also rises and the velocity eventually falls back to point “c.” This new, steady
operating point represents a substantial reduction in slurry throughput and corre-
sponds to point “c” in the previous figure. However, if the low-concentration slurry
is now introduced into the pump, the pump characteristic falls and the new intersec-
tion point falls to zero velocity at point “d.” At this point, the pump is only producing
enough pressure to balance the yield stress of the high-concentration slurry in the
pipeline, and the system flow stops. Introducing flush water into the pump only
aggravates the condition, as the pump characteristic now falls even lower. The
system can only be restarted by introducing a higher concentration slurry into the
pump, in order to bring the pump characteristic somewhat above the
low-concentration value, or by increasing pump speed. This presents another exam-
ple where variable pump speed capability provides critical support for maintaining
system stability.
Now consider the case with variable-speed pump operation controlled to maintain
constant slurry throughput. The first effect of the increase in solids concentration will
be to reduce pump speed, to prevent “overshoot” towards point “b,” and to maintain
operation at point “a.” As the concentration front moves through the pipe and the
system characteristic rises to the high-concentration curve, the pump must then
increase speed to maintain an intersection between the rising system curve and the
pump curve at the same velocity as point “a,” albeit at a higher system head.
Eventually, point “e” is reached when the high concentration propagates fully
through the system. Note that the pump characteristic (dashed line) lies above the
fixed speed, high-concentration pump characteristic, indicating that the speed has
exceeded the starting value. A similar but reversed process can be followed due to a
concentration decrease.
Note that Figs. 10.5 and 10.6 refer only to the effects of changes in solids
concentration. Changes in slurry rheology due to variations in particle size may
alter the system curves in a fashion similar to the concentration increase, but the
pump curves may respond differently since the concentration (and therefore the
slurry density) is not changing. As a result, the effects of “overshoot” and recovery
will change; however, operation can still vary, and the approach outlined above can
be followed to determine the actual effects and develop solutions. Non-Newtonian
slurries with coarse solids that may settle and form a bed (i.e., complex slurries) will
present further complexity, as the turbulent curves will no longer converge on the jm
plot and variability may increase further. In many cases, the best solutions will
involve the application of a variable-speed pump driver, which can then be factored
into the analysis.
Because of the complex nature of slurry flows, problems in operating slurry
pipeline systems can arise from a variety of causes. The sources of unstable or
10.6 Specific Energy Consumption in System Design 313
inoperable conditions in this chapter have been deliberately simplified, with only one
variable considered at a time to illustrate the principles involved. In actual practice,
the variables and dynamics are often more varied, and the reasons for problems
encountered in operating slurry systems are sometimes far from obvious. However,
by the methodical application of the principles outlined above, most cases can be
modeled, diagnosed, and corrected, or (ideally) anticipated and factored into the
slurry system design from the beginning.
250
D = 300mm
Pressure gradient (kPa/m)
300
Cv Ss = 2.65 350
t = e10 cv +0.01(1+cv) g
400450
turbulent 500
laminar
(a) 250 mm
300
0 2 4 6
350
Velocity (m/s)
400
(a) 450
500
Fig. 10.7 Variation of SEC with concentration and pipe size for a homogeneous, non-Newtonian
slurry. (a) Transport characteristics showing 600 tph and 600 dtph contours, (b) SEC for the mixture
transport as a function of Cv and pipe diameter, (c) SEC for solids transport only
flow velocity for various delivered concentrations. Contours for 600 tph (ton per
hour) of the mixture and 600 of the dry solids (dtph) have been marked, assuming
these could be separated from the slurry in some way. Figure 10.7b shows the
variation in SEC with solids concentration and a variety of pipe sizes for the fixed
tonnage of mixture, while Fig. 10.7c shows similar information for the fixed tonnage
of dry solids.
For the fixed mixture tonnage case (Fig. 10.7b), the flow regime transitions from
laminar to turbulent flow (within any pipe size) as the concentration decreases and
the velocity increases to maintain a fixed throughput. For each pipe size, the
minimum SEC is found near the transition point from laminar to turbulent flow,
shown by the dashed line. To the left of each minimum, where the flow is turbulent,
decreasing the solids concentration increases the SEC, as predicted by the equations
in Table 10.1, whereas to the right of the minimum, the reverse is true. This is
because the flow has become laminar and is dominated by the viscosity of the slurry,
which is a strong function of solids concentration. Increasing the concentration at
these levels produces a much more viscous slurry that requires a higher transport
pressure gradient, and consequently a higher SEC. The fixed dry solids tonnage case
shown in Fig. 10.7c has similar characteristics to Fig. 10.7b, but the SEC values are
much higher, especially at the lower concentrations, since no value is placed on the
underlying fluid.
10.6 Specific Energy Consumption in System Design 315
D = 300mm
Ss = 2.65
200mm
Pressure gradient (kPa/m)
350
40% 400
30%
20%
10%
Ms = 700 dtph
Ss = 2.65
Fig. 10.8 Variation of SEC with solids concentration and pipe size for a heterogeneous settling
slurry. (a) Pipe friction characteristic curves for a 300 mm pipe including a 700 dtph contour.
(b) Variation of SEC with Cvd and pipe diameter for a constant tonnage of 700 dtph
11.1 Introduction
Earlier chapters in this book were concerned with the basic principles of design,
selection, application, and operability. This chapter is devoted to some of the
practical issues and problems that can arise during operation in the field, such as
the transient effects due to speed change, bed deposition, water hammer and reverse
flow. Multi-pump systems are discussed, both in series and parallel, including start-
up and shutdown protocols. Sump and suction piping design, as well as the pumping
of frothy mixtures, are discussed.
Optimization of slurry transport systems must often include practical consider-
ations outside of the technical and theoretical realm, such as the scarcity of capital
and the various standards, regulations, geographical and cultural constraints facing
any given installation. Even the relative importance of the various technical consid-
erations is not always easily weighed without the perspective given by actual
operating experience. Problems in operation may arise when some element of the
system—often the pump or drive—is not correctly selected or installed, perhaps
because it was on-hand rather than being specifically chosen for the duty. Other
culprits include faulty design assumptions, poor operating practice, misleading
instrumentation, and incorrect analysis of data, any of which may not be immedi-
ately apparent. Finally, in the realm of slurry transport, one must learn to expect the
unexpected, and problems are sometimes encountered without precedent that require
new solutions not previously employed.
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 317
R. Visintainer et al., Slurry Transport Using Centrifugal Pumps,
https://doi.org/10.1007/978-3-031-25440-6_11
318 11 Practical Experience with Slurry Systems
Considerations of wear and ease of maintenance usually dictate that the centrifugal
slurry pump be a single-stage machine. Furthermore, because of limitations on the
turning speed of the impeller dictated by strength of materials and/or wear consid-
erations, a single centrifugal pump cannot be used to transport slurry over a distance
of more than a kilometer or so. Multiple pumps in series may therefore be needed for
medium- or long-distance pipelines. Pipelines as long as 25 km and using up to
17 pumps in series are currently in service.
The total head generated at any flow rate by a collection of pumps in series is
simply the sum of the heads developed by each individual pump at the common flow
rate. Several alternative selections are often considered for any given system, based
on different numbers of pumps in series. As head per pump increases, the number of
pumps required decreases, as does the hydraulic size of the pump (assuming a fixed
flow rate); however, the pump speed and wear rate increase. In many cases, the
lowest capital cost is achieved by applying a smaller number of pumps at higher head
per stage, while the lowest operating cost, including parts, maintenance, and down-
time, is achieved by applying more pumps at lower head. For larger systems, which
are costly to build and operate, a total cost of ownership calculation should be made,
taking into account the costs of capital, power, replacement parts, maintenance, and
downtime per unit mass of solids delivered (see Chap. 14).
For longer pipelines and finer solids, the alternative of a positive displacement
pump may be considered. Although these often have higher capital and maintenance
costs, they can handle higher heads per stage, reducing the number of machines
needed, and will have a better efficiency when the slurry viscosity is high. As a
result, positive displacement pumps are viable alternatives to centrifugal slurry
pumps in systems that transport highly non-Newtonian slurries over long distances.
When operating multiple centrifugal pumps in series, it is generally preferred that
all pumps be of the same hydraulic size. This simplifies the operational control
philosophy and helps ensure that the pumps experience a similar maintenance cycle.
However, it is often necessary to have a few variable speed pumps in the line. This
helps control pipeline flow rate as the hydraulic gradient varies due to changes in the
concentration and/or particle size distribution of the slurry, or as pump performance
varies due to wear. Without these pumps, a long pipeline may be difficult to control,
which can lead to various problems such as pump cavitation (flow too high),
plugging of the line (flow too low), or other undesirable transient conditions that
have a negative impact on the slurry process and wear of the pumps and pipeline
components. As variable-speed controllers for induction motors have become more
reliable and cost-effective, many long-distance pipelines have been converted to use
variable-speed drivers on all pumps. This allows all pumps to operate at similar
speed with similar loading, avoiding the differences that occur when pumps are
operated at different speeds with a fixed flow rate. Such speed matching ensures the
most even distribution of wear between pumps and the longest overall service life. It
also allows for greater ease of operation, especially during start-up and shutdown.
11.2 Pumps in Series 319
Spacing the pumps at roughly equal hydraulic distances along the pipeline will help
limit the pressure each pump must withstand. This configuration, however, can
increase the difficulties associated with start-up and control, and add to the cost of
providing foundations, power, and other facilities to the widely spaced pumps. Care
must also be taken to ensure that each pump in the line receives sufficient suction
pressure to avoid cavitation at all operating conditions. Conversely, if all pumps are
grouped together, the facilities costs are reduced and the suction conditions simpli-
fied, but the downstream pumps will experience elevated operating pressures in
proportion to the number of pumps grouped together. The result is a higher cost of
construction for the pump casings and shaft seals. In many cases, a hybrid approach
is used, spacing groups of two or three pumps together at each location.
In a long pipeline, care must be taken in the proper sequence and timing of pump
start-up, and variable-speed drives may again be required on one or two of the pumps
in the system for control purposes. The start-up sequence and timing of a multiple-
pump system must be considered in light of the effect on the hydraulic grade line,
and how this effect varies with time. If all pumps are started simultaneously at full
speed in an already filled pipeline, a brief initial period will follow in which the
pipeline flow, and hence the frictional head loss, are negligibly small. In a horizontal
pipeline, this implies that the downstream pumps will, at least for a short time, be
subjected to the combined zero-flow pressure generated by the upstream pumps,
even if they have been spaced out along the line. Figure 11.1 illustrates this dynamic
for one-, two-, and four-pump systems. The start-up pressure for such a system can
be much more severe than that of the normal steady-state pressure. In some recorded
cases, multiple pumps have experienced catastrophic casing rupture because of this
initial over-pressure.
To avoid overstressing the pipeline (and the pumps), pumps spaced along a
pipeline must be started in a timed order arranged to minimize the maximum
pressure. In general, the start-up sequence for a multi-pump system should begin
with the first pump in the line, progressing to each downstream pump in turn after a
suitable delay, to allow flow to begin. This approach avoids creating low suction
pressure and cavitation at any single pump while allowing flow and hydraulic
gradient to build as more pumps are brought online and system pressures increase.
While this approach holds as a general rule and represents the “safe option,” note that
for longer lines, more complex sequences (such as 1, 3, 2, 4) may be more effective
(i.e., both faster and resulting in a smaller peak pressure) when pumps are evenly
spaced along a horizontal line. Computer simulation of the transient flow conditions
during start-up is recommended for these cases. Also important is the issue of
“pumping through” those idle pumps that have not yet started (see Sect. 11.2.4).
320 11 Practical Experience with Slurry Systems
B A
1 Pu m p
B D
A
2 Pumps
C
Pump 1
Pump 2
B
4 Pumps
A D
Head (Pressure) >>>
C
Pump 1
Pump 2
Pump 3
Pump 4
The above examples assume a level pipeline with no changes in elevation profile;
however, elevation changes are common and must be considered for their impact on
system head and the hydraulic grade line, and also for their effect on system start-up,
shutdown, and transient operations. For example, when the line is shutdown, any
local high spots will be susceptible to subatmospheric pressures. If large enough,
these pressures will cause the liquid to vaporize. On subsequent start-up, collapse of
the vapor pocket can initiate severe hydraulic transients. To avoid these issues,
vacuum relief valves should be fitted at high points, or else the line must be
completely drained after each shutdown.
High points before long downhill runs may see subatmospheric pressures even
during normal operation, if the head gained by the elevation drop exceeds that of the
slurry friction loss within the pipe. Again, a vacuum relief valve placed at the high
point can be used to prevent vaporization, but this valve will also admit air. In both
cases, the flow in the downhill section will convert to open-channel or slack flow,
increasing in velocity and possibly causing problems with pipeline wear or opera-
tional stability. The pipe section in open-channel flow will not experience head
recovery due to the change in elevation, and a great deal of turbulence will occur
where the pipe returns to the fully filled condition and the higher velocity of the open
channel flow is arrested.
To avoid these problems, long downhill runs are sometimes fitted with choke
stations at the bottom of the run. These are sections of pipe that produce additional
system resistance by using smaller pipes, multiple elbows, orifices, or some combi-
nation of these, such that the pressure at the high point remains above the vapor
pressure and the pipeline remains full. (Betinol and Jaime 2004; Derammelaere and
Shou 2002; Rusconi et al. 2016). Similar chokes may be required when placing
backfill in worked-out areas of deep mines (Wilkens et al. 2004). Although the choke
station components will experience wear, increasing maintenance costs, these may
be justified by the gain in in system stability.
Another alternative for this case is the placement of slurry pumps in the downhill
line, operating in reverse rotation as turbines to absorb the energy and convert it to
power by means of attached generators (Brewster 1993; Cave 1980). Although this
is an appealing solution from an energy conservation standpoint, in practice the extra
cost and energy expenditure required to build and maintain the slurry turbine stations
makes this alternative less desirable than it might first appear, and installations to
date have been limited.
In locating pumps along the line, the system net positive suction head available at
each pump (the NPSHA) must be considered and prevented from falling below that
required by each pump (the NPSHR) for all desired operating conditions, including
322 11 Practical Experience with Slurry Systems
the transient conditions experienced during start-up and shutdown. Cavitation of one
pump in a long pipeline can affect the entire train, in the worst case leading to water
hammer, or a loss of flow and pipeline plugging. For safety, it is desirable to
maintain a minimum of one atmosphere of positive pressure at the suction of each
pump. In some cases, the line may even be broken, with an open sump at every
pumping station, to prevent hydraulic transients from being transmitted down the
line; however, this approach increases operating costs by destroying head and can
lead to other problems associated with sump level control. Tanks may also be
employed at high points in the line to maintain atmospheric pressure and prevent
slack (open-channel) flow.
In some cases, the first pump at the suction end of the pipeline is run at lower
speed than the other pumps in order to reduce its NPSHR and allow a system design
with lower NPSHA at that location. This setup is most common in dredging systems
where a submerged ladder pump operating at lower speed is deployed to improve the
vessel’s suction characteristics while simultaneously providing a high NPSHA to the
on-board pump, which may then produce more head by running at higher speed. For
convenience in stocking spares, the ladder and on-board pumps are sometimes made
identical, but they may also be of different sizes. In particular, the use of a higher
specific speed pump on the ladder can reduce its size (and its driver size) while still
providing a good hydraulic match to the larger on-board pump.
Figure 11.2 illustrates the case of identical pumps in series driven at different
speeds. In this example, the speed of the ladder pump is two-thirds that of the
on-board pump (400 rpm and 600 rpm, respectively). As always for pumps in series,
the total head developed at any flow rate is the sum of the individual pump heads,
leading to the combined head curve shown. Because of the ladder pump’s lower
System A
System B
System C
2
Q1 Q2 Q3 Flowrate >>>
Fig. 11.2 Pump and system characteristics for a dredge with ladder pump operating at reduced
speed
11.2 Pumps in Series 323
speed, and hence its smaller QBEP, its NPSHR is lower than that of the on-board
pump at flow rates up to about 133% of the ladder pump’s QBEP, after which the
advantage is lost. The difference in the location of QBEP between pumps also leads to
differences in operating efficiency at most flow rates.
Let us now consider various operating conditions for this two-pump system. For
System A, characterized by high resistance, the combined head curve intersects the
system curve at point 1, and the operating flow rate is Q1. The ladder pump is
operating at 120% of QBEP, with some reduced efficiency relative to the maximum,
while the on-board pump operates at 80% of QBEP, at roughly the same reduced
efficiency. At this flow rate, the lower speed ladder pump has a lower NPSHR, even
though it is operating at a higher percentage of QBEP.
Consider now the case where the system resistance is reduced, for example by
pumping through a shorter line, as represented by System B. The combined head
curve now intersects the system curve at point 2, and the operating flow rate is Q2.
With a shorter pipeline and lower system head, this might appear to be an easier
duty. However, as is sometimes said of amatory porcupines, care is needed in
addressing the task. While the on-board pump is operating at 100% of QBEP and
optimum efficiency, the ladder pump is now operating near 150% of QBEP, where
efficiency is greatly reduced and the NPSHR exceeds that required by the higher
speed booster pump.
Finally, with a further reduction in pipeline resistance to System C, operation
shifts to point 3 and the flow rate is increased to Q3. At this flow rate, the low speed
ladder pump has reached the point of zero head and no longer contributes to the
combined head curve. Further reductions in system resistance, along with the
corresponding increases in flow rate, would produce a condition where the low
speed pump actually absorbs head. (It is assumed that sufficient NPSH is available to
prevent cavitation, perhaps because the ladder pump is immersed more deeply.)
When this condition is encountered in the field, it is sometimes interpreted as the
result of cavitation or poor performance of the ladder pump, when it is, in fact, a
simple outcome of the mismatch in pump speeds.
This example illustrates the importance of checking system operating points and
suction conditions, especially when using series pumps that are not identical or are
not driven at the same speed. It also shows that in some circumstances, the NPSH
required by a pump may not be lower just because it is driven at a lower speed.
A similar analysis can be made for the case of two pumps that differ in hydraulic
size or specific speed. The complexity of the analysis will increase somewhat, but the
same approach can be applied, and the additional complexity may be justified by an
increased suction performance or other operating characteristic of the pair.
Because of the higher pressures involved, multiple-unit pumping stations most often
use bases bolted to concrete foundations. The unbalanced forces due to pressure
across the suction and discharge of the pump can become quite large, and some
determination must be made regarding where these forces will be carried (i.e.,
11.2 Pumps in Series 325
Net force of
unbalanced
pressure load
Hydraulic load
Portion of load
carried into piping
Portion of load
carried into pump base
through the pump into the foundation, or in tension back into the piping itself, or
some combination of these two extremes, as illustrated in Fig. 11.3). In principle,
assuming a very rigid pump with flexible piping (e.g., an expansion joint), these
loads are carried through the pump and into the base. In the opposite extreme of a
completely rigid pipe, the loads are carried back into the piping itself. In practice,
neither extreme is easily reached. Most industrial installations will fall somewhere
between.
In addition to the hydraulic load, other mechanical loads, such as those due to
pipe expansion with temperature, must be considered. In addition, any increase in
the pipe diameter between the pump and any flexible element, such as an expansion
joint, will result in an additional load on the pump, equivalent to the pipeline
pressure multiplied by the added area of the larger pipe, as illustrated in Fig. 11.4.
Such loads can become quite large, and the temptation to use the pump as a “piping
anchor” in order to carry them should be discouraged. The centrifugal slurry pump is
one of the most expensive components in a slurry pipeline system, from both a
capital and an operational point of view, and is often made of exotic and sometimes
brittle materials subject to wear. The practice of applying unnecessary loads to the
pumps can end up costing much more in operational problems than the up-front cost
of a more robust pipeline anchoring system that is independent of the pumps
themselves. For critical systems, current practice is to analyze both the pump and
piping using FEM methods that account for the stiffness of all components and
provide the forces to be carried into the pump and piping anchors, as well as the
internal stresses within the components themselves.
The layout of pumps within a series station depends on the number of units and
the requirements of maintenance, access, and safety. Figure 11.5 shows a photo-
graph of a six-pump station with two lines of three pumps in series. In this case,
access to all pumps is excellent, and reasonable work areas are provided adjacent to
326 11 Practical Experience with Slurry Systems
Fig. 11.5 Interior of pumping station showing two sets of three pumps in series
11.3 Pumps in Parallel 327
Fig. 11.6 Interior of pumping station showing compact arrangement of pumps in series
each pump. Note that the use of different levels makes it possible to employ single-
bend pipe sections between pumps. More compact arrangements can be achieved by
rotating the pump branches and using straight pipes between each adjacent discharge
and suction, as seen in Fig. 11.6.
Fig. 11.7 Pump and system characteristics for three pumps operating in parallel
system head curve. Because slurry pump head curves are often “flat” (i.e., not having
a steep slope relative to flow rate), the individual pump performances in parallel
systems are subject to larger variations relative to each other than they are in series
systems, and systems with pumps operating in parallel are more sensitive to prob-
lems arising from improper sizing or from difficulties in balancing and control. To
maintain practical control of parallel systems, it is usually required that all the pumps
be hydraulically identical, even to the point of being in the same state of repair, since
a worn pump may have a different performance characteristic than one that has been
newly rebuilt.
Figure 11.7 illustrates the case of three identical pumps running at the same speed
and discharging in parallel to a common header. Parallel systems of this type are
sometimes installed to allow variations in system flow rate by varying the number of
fixed-speed pumps in operation. With all three pumps in operation, the total system
flow is Q3. Each individual pump operates at flow Q13, with power, efficiency, and
NPSHR corresponding to the single pump curves shown. Note that flow Q13 happens
to be near QBEP for the individual pump. If one pump is taken out of service, the total
system flow is reduced to Q2, but each of the two remaining pumps moves out to
flow rate Q12. Note that while the overall system flow rate has decreased, the
remaining pumps are now operating at a higher individual flow rate and require
more power and NPSHR than when all three pumps were in service. Also, because
they are now operating beyond QBEP, their efficiency has decreased, further increas-
ing the power requirement. This may seem counterintuitive at first, but it is the
natural result of fewer pumps competing for the same fixed system capacity. With
only one pump in service, the system and pump operate at point Q11 with a further
increase in required pump power and NPSHR, and a further reduction in efficiency.
11.4 Start-up, Shutdown, and Transient Conditions 329
In designing a parallel system such as that shown in Fig. 11.7, the pumps and
drives must be selected to operate at all three conditions—Q11, Q12, and Q13—with
the highest flow rate, power, and NPSHR typically seen at condition Q11. For best
overall operating efficiency, the pumps should be sized such that the QBEP is located
near the most common condition; however, it should also be considered that placing
the QBEP closer to condition Q11 will lower the maximum required power and may
thereby reduce the size and power rating of the drivers.
The variations in pump power and torque during start-up depend on variations in the
hydraulic gradient of the system in which it runs, and in particular, the relative
proportions of friction and static head. Also important is whether the line is full or
empty at start-up. Consider the three system resistance curves A, B, and C shown in
Fig. 11.8, representing systems with a low, medium, and high proportion of static
head, respectively, but with all having the same normal operation condition at head
H1 and flow rate Q1. All systems are assumed to be filled with fluid upon start-up.
For System C, the pump is required to deliver to a pipe with a large static lift, and
flow can begin only after the pump has reached speed Y, where it begins to produce
enough head to overcome the static requirement. In other words, the flow in this
system will start only after this speed is reached. For System A, by comparison, flow
begins almost as soon as the pump begins turning. System B presents an
H1
H0,C
Head
H0,B
H0,A
Flowrate Q1
Normal
100% Operating
Point
% Normal Operating Torque
T0
100%
% Normal Operating Speed
Fig. 11.9 Variation of torque with speed for pump starting (system initially full)
intermediate case. In all three systems, the same condition of head and flow is
reached at pump speed Z.
Figure 11.9 shows the pump torque-versus-speed curves corresponding to the
system variations given in Fig. 11.8. All three cases begin in a similar way: At low
speed, where mechanical friction in the bearings and seals dominates. A starting
torque (T0) in the range of 10–20% of full running torque is required to get the pump
moving. As the static friction is overcome and the pump begins to turn, the lower
dynamic friction (still mechanical) results in a dip in the torque-speed curve. The
curve eventually increases again with speed, as acceleration of the flow in the pump
takes over as the main driver of pump torque. In the case of System A, where system
flow begins almost immediately, the curve follows a roughly parabolic shape up to
the operating point at 100% speed and 100% torque, assuming a typical low-to-
medium specific speed slurry pump is being used. In the case of System C, no flow
begins until pump speed Y is reached; therefore, the torque requirement is lower,
being limited to that needed to overcome mechanical friction and the recirculation of
fluid within the pump. Once pump speed Y is reached, the flow and torque increase
quickly until the same operating condition is reached. Intermediate cases such as
System B will fall within the shaded area between curves.
Although not related to pump start-up, it should be noted that the flow rate for
System C, with a higher proportion of static head, is much more sensitive to pump
speed than the low static head System A, achieving the full range of flow with a
much smaller range of speed. Such systems are common in slurry process plant
applications. The sensitivity of these systems to speed changes—or to other factors
affecting the system curve itself, such as changing sump levels or slurry
11.4 Start-up, Shutdown, and Transient Conditions 331
Fig. 11.10 Stabilizing influence of sump level on variations in system flow rate
H1
Head
Flowrate Q1 Q2
shown in Fig. 11.11, and leads to a surge in the required driver power due to both the
increased flow rate and slurry density. This increased density also decreases NSPHA,
and if the flow increase is large enough, there is an additional risk of cavitation and
water hammer. A sudden reduction in slurry concentration causes a reduction in
flow, such that if the system is operating close to the deposition limit velocity, the
solids can settle out. The resultant problems in restarting the line were discussed in
Chap. 10.
In many cases, the proper approach for a controlled shutdown is simply a reverse
of the start-up protocol. Additional considerations may include the need to flush the
solids from the system before the shutdown, and provisions for potential backflow or
draining of the pipeline in systems with significant static head (see Sect. 11.7).
The conditions resulting from an unplanned shutdown should also be examined.
There are many different potential causes for unplanned shutdowns, and each may
have its own particular risks and challenges. Common unplanned events include:
• Power failure
• Pump trip due to motor overload, excessive vibration levels, oil temperature, or
other emergencies
• Pump blockage due to excessive solids or tramp material
• Loss of upstream process feed (solids and/or liquid)
• Breakdown of upstream or downstream process equipment
The risk factors resulting from an unplanned shutdown vary according to the
details of each system design and operating plan. Common factors encountered
during a typical risk assessment include:
• The deposition of solids in the pipeline, sump, and process equipment, and their
potential for clogging and preventing a restart
• Reverse flow and impeller unscrewing
• Pump cavitation and risk of water hammer due to sudden deceleration of pump
speed, especially if operating at a high pipeline velocity
• Impact on the upstream and downstream process equipment
Other transient conditions may occur during normal operation due to changes in
the concentration or particle size of the feed, pump cavitation or blockages, varia-
tions in sump level, and other causes. These are addressed throughout the text,
especially in Chap. 10. All potential transients and associated operating protocols
should be considered when designing any slurry system.
As noted in earlier chapters, operation near or below the limit of stationary deposi-
tion may lead to instability in systems using centrifugal pumps and is therefore
normally avoided. Accelerated pipe wear may also result from the secondary flows
and corrosive effects that can occur at the interface between a stationary bed and the
334 11 Practical Experience with Slurry Systems
pipe wall. For a fixed throughput of solids, the operating velocity is ideally kept
above the deposition limit velocity, and the design is optimized by careful selection
of the pipe diameter and solids concentration.
However, some systems may experience a widely variable solids throughput due
to the conditions of the process, and it may be necessary to operate, at least part of
the time, with a stationary deposit. In analyzing systems of this type, two key factors
must be considered: the stability of the deposition event (when it occurs), and the
resulting friction loss in the partially filled pipe.
A first approximation of the friction loss of a settling slurry operating with a
stationary bed can be made by replacing the mixture velocity and diameter in the
appropriate friction loss model (Chap. 5) with the actual velocity above the bed and
the hydraulic diameter of the partially filled pipe. It may be necessary to carry out a
trial-and-error iteration of this calculation to determine the depth of the stationary
deposit. The correct depth will be that where the velocity of the partly filled pipeline
is equal to the deposition limit velocity of the worst-case particle size in the slurry.
This approximate approach will yield less accurate estimates as the bed depth grows,
due to the greater deviation of the actual flow channel shape from the assumed
circular pipe, and also due to the roughness of the bed itself. Both will tend to further
increase friction loss. However, the approach is often adequate for preliminary
design and troubleshooting. As discussed in Sect. 5.6.1, a more accurate and
complete calculation can be made by using a unified layered model.
Regarding stability, the most stable transition from suspension to deposition
occurs with low concentration mixtures of fine settling solids—for example, a slurry
mixture with d50 = 200 μm and Cv = 15%. As the velocity is decreased, deposition
occurs gradually until an equilibrium between the bed depth and mixture velocity is
reached. The opposite occurs as the velocity is increased and the deposit is
resuspended. At the other end of the spectrum, high-concentration mixtures of solids
in the heterogeneous range (200 μm < d < 0.015D) can be especially problematic—
for example, a slurry mixture in a 200 mm pipe with d50 = 2 mm and Cv = 35%.
Such mixtures tend to deposit suddenly as a complete bed and are also more subject
to the formation of density waves along the pipe when operating near the deposition
limit velocity (see Sect. 5.6.2). High-concentration mixtures of fully stratified solids
(d > 0.015D) are also problematic for similar reasons. Mitigating factors include the
speed with which the transition to deposition is made and the absence of disconti-
nuities in the pipeline (elbows, low spots, etc.) that may serve to collect solids and
lead to a plug during deposition. The safest condition will be one where the pipeline
is relatively horizontal and straight, and the velocity reduction occurs gradually
allowing time for the deposit to stabilize.
The above discussion addresses settling slurries, but similar conditions may arise
in non-Newtonian slurries, although the dynamics are considerably more complex.
This is especially true when coarse solids are being transported in a non-Newtonian
carrier. Observation of such flows can be difficult, as they often remain opaque
during deposition, and simple determination of the deposition limit velocity alone
can be challenging. Reliance on previous operating experience or testing is often
recommended (see Chap. 6).
11.6 Water Hammer 335
The term water hammer is applied to elastic pressure transients in a liquid or slurry
that are caused by a sudden stop or change in direction of the fluid motion. It most
commonly occurs in pipelines due to sudden valve closures or collapsing vapor
pockets. The literature on such pressure transients is extensive; for more details, the
reader is referred to the classic text of Parmakian (1955) or to subsequent works by
Wylie and Streeter (1978), Sharp (1981), and Chaudhry (1987).
When generated, an elastic pressure transient or “pressure wave” moves through
the fluid at a characteristic sonic velocity or celerity (a), which is related to the speed
of sound in the fluid or slurry mixture. In a pipeline, the elasticity of the walls
reduces the celerity to a value somewhat below the speed of sound in that mixture;
however, for a steel pipe filled with water (or an aqueous slurry), it is still about 1000
m/s. As given by the elastic-column model and verified experimentally, the rise in
pressure generated by a change in velocity is:
Δp = - a ρf Δ V ð11:1Þ
where ΔV is the change in velocity. In terms of head, the transient rise is:
a
ΔH = - ΔV ð11:2Þ
g
Because the ratio a/g is approximately 100, the head transient in a pipeline is
about 100 times the velocity change.
Equations (11.1) and (11.2) apply to any rapid change in operating conditions.
For example, rapid closure of a valve in a pipe causes the pressure to rise on the
upstream side and fall on the downstream side. If the pressure on the downstream
side falls below the vapor pressure of the fluid, a vapor cavity will form. Equation
(11.1) shows that this happens for quite modest flow velocities. Subsequent collapse
of the vapor cavity causes the valve to experience a water hammer in the reverse
direction as it is struck by the column of fluid moving back to fill the cavity. This
sequence is illustrated in Fig. 11.12.
Rapid opening and closing of valves should always be avoided. However, vapor
pockets can also be caused by other types of maloperation, of which the most
common is sudden gross cavitation in a pump. (See Sect. 8.3 for more information
about cavitation and the conditions that cause it.) Carstens and Hagler (1964)
showed how vapor pockets caused by poor pump operation can lead to severe
water hammer. The resulting pressure transients can be quite large. For example,
vapor pocket closure at the relatively modest velocity of 5 m/s (16.4 ft/s) causes a
pressure peak of about 50 bar (i.e., 500 m or 1640 feet of head).
It is important to understand that the peak thus generated is not just a local event;
rather, the initial disturbance initiates pressure waves that travel in either direction
along the pipe at velocity a. In a pipe of constant cross-section, pressure waves pass
by each other without changing amplitude or waveform. However, when they reach
336 11 Practical Experience with Slurry Systems
the end of the pipe or a change in section, the pressure waves are reflected. The
maximum pressure transients thus arise at points of initiation and reflection, or where
different wave fronts meet. As a result, maloperation of the pump at the feed end of a
long line can set up a pressure wave that travels along the line until it meets the first
change in section. This is often the next pump, which may suffer more or less severe
damage. In extreme cases, the water hammer may fracture the pump casing, or the
forces generated by the passing pressure wave may move the pump on its base,
disrupting the alignment of the drive train and causing further damage to the drive
components and supporting structures.
Similar behavior occurs in pipelines containing slurries, although here the accel-
eration of the particles modifies the pressure rise, decreasing it for most particle
materials but also possibly increasing it for materials like coal (Wood and Kao 1966;
Han et al. 1998; Cristoffanini et al. 2014; Kodura et al. 2019). Minor air entrainment,
usually present in slurry flows, has a suppressing effect through its compressibility,
as do compliant pipeline linings, which often reduce the pressure transient ampli-
tudes for these flows.
11.7 Reverse Flow 337
Upon the shutdown of a system with significant static lift, the fluid remaining in the
pipeline may flow backward through the de-energized pumps, causing them to act as
turbines and rotate in reverse. The primary dangers are over-speed and the potential
for negative shaft torque under certain conditions. Since many slurry pump impellers
are fastened to their shafts by a threaded connection, negative torque can lead to
unscrewing of the impeller and the potential for extensive damage to the pump and
drive train.
The following provides a summary of the sequence of events for reverse flow in a
system with a significant static head driven by a typical centrifugal slurry pump. The
example is given for one pump, but the sequence remains essentially the same for
multiple pumps, although their spacing along the line may complicate the scenario.
1. Shortly after the pump is de-energized, it slows to a speed where the static head of
fluid in the system is balanced by the zero-flow head produced by the pump (i.e.,
its shut head). At this point, the flow has dropped to zero, but the pump is still
turning in the forward direction, carried by the inertia of the impeller at 60–80%
of full speed.
2. Reverse flow now begins, and the rotational speed of the pump (still in the
positive direction) continues to slow.
3. At the point where the pump speed reaches zero, the reverse flow will be
approaching its maximum value.
4. The pump impeller now begins to accelerate into reverse rotation, driven as a
turbine by the reverse flow. Since the pump is de-energized, the only resistance to
rotation will be friction in the bearings, seals, couplings, and gears, so the reverse
speed may exceed normal forward speed.
5. As the system empties and the static head reduces, the pump slows and comes to a
stop as the flow stops.
If the sequence of events as described above is allowed to proceed undisturbed,
the torque on the pump shaft remains positive and the impeller will not unscrew.
Assuming that the various elements of the drive train are rated for reverse rotation
and will remain lubricated during the time required for the pump to come to a stop,
the system can be allowed to drain naturally, without intervention.
However, for a long system, this sequence may take a long time, as much as
30 min or more. Also, the amount of fluid draining back may cause problems if there
is no convenient place for it to be collected. In these cases, some attempts may be
made to divert the flow and shorten the time needed to reset the pumps and restart the
system. Care must be taken when this is done. If the reverse flow through a pump is
reduced faster than the inertia of the impeller will allow its rotation to reduce, then
the fluid in the system will drag, rather than drive, the impeller, and the resulting
torque will be negative. As a rule of thumb, a valve closure or flow diversion time of
three minutes will provide ample time for the impeller to slow without unscrewing.
Shorter times may be possible based on a detailed transient flow analysis of the
particular system.
338 11 Practical Experience with Slurry Systems
One question is often asked whether the system can be valved off or diverted just
at the point where the flow reaches zero, preventing backflow altogether. This is
easily accomplished by the installation of a check valve near the bottom of the
system, a solution often applied in clear fluid systems. However, this solution is not
easily implemented in most slurry systems, where the slurry’s abrasive nature would
damage and incapacitate most check valves during normal operation. Theoretically,
the same end could be accomplished by a carefully timed valve closure at the point
where the flow reaches zero; however, in practice, this is virtually impossible to
manage in a reliable way, as the time required for the flow to reverse is usually very
short—often on the order of 10 s or less—and the window of time where the flow is
near zero may last for no more than a second or two. A poorly timed valve closure
could lead to water hammer, the impeller unscrewing, or both.
Another possibility is a clutch in the drive train designed to prevent reverse
rotation. This will not prevent reverse flow, but by locking the rotor against reverse
rotation, a positive torque will be maintained as the fluid drains. While such devices
exist, they are usually limited to smaller pumps and become increasingly expensive
and bulky as the pump size increases. As a result, they are not often used in practice.
In addition to the above considerations, there are several special cases which bear
closer examination:
• Shutdown due to power failure. In this case, the auxiliary systems for drive train
lubrication and pump seal water may be disabled. Depending on the length of
time required for the backflow, some provision for maintaining these services
may be required.
• High static head. Systems with high proportions of static head (>85%), espe-
cially those with multiple pumps in series at the bottom of the slope, contain a lot
of energy and may achieve reverse flows and speeds well in excess of the normal
forward values—in some cases, up to 50% greater. The structural integrity of the
piping and piping supports, as well as the drive train components (bearings, seals,
couplings, gears, and impeller) must be checked to ensure that they are rated for
this condition, including the effects of torsional or lateral resonance at the higher
turning frequencies.
• Short, mostly vertical system. In this case, the pipeline may empty while the pump
is still turning at a high reverse speed. If fluid remains in the pump after the
pipeline empties (perhaps due to the adjacent sump level), a condition similar to
the diversion of flow mentioned above occurs: the fluid in the pump drags on the
impeller, and the torque becomes negative. These systems are also subject to
clogging if coarse and high-concentration solids are being pumped and the system
shutdown occurs before the solids can be flushed out.
• Operation with torsional natural frequency. Pumps are sometimes operated with
a torsional natural frequency in the operating range that is avoided (passed
through quickly) during normal operation. During de-energized reverse rotation,
this can lead to the loosening of the impeller as the natural frequency is passed
during slowdown and reverse rotation. Therefore, slurry pumps with threaded
impellers subject to reverse flow should not be allowed to operate through a
torsional natural frequency.
11.8 Pump Explosion 339
A topic of particular concern for the safe operation of all centrifugal pumps is the
potential for pump explosion due to simultaneous blockage of the pump suction and
discharge. If a centrifugal pump is run with such a blockage, there will be no flow
rate, and the power expended in turning the pump will heat the fluid within the
pump. If this heat cannot be dissipated by conduction through the pump casing faster
than it is being generated, the fluid temperature will rise. If the blockage is pressure
tight, the pressure within the pump will also rise, and if allowed to continue
unchecked, can exceed the burst pressure of the pump casing. Unlike a cold water
hydrostatic pressure test, or even a water hammer pressure surge, a casing fracture
due to overpressure by superheated water releases a great deal of energy as the water
vaporizes and expands. This event can be quite violent, depending on the conditions.
In clear fluid centrifugal pump applications, the most common cause of such
events is an accidental closure of valves, isolating both the suction and discharge of
the pump. In high head, multistage pump applications, the overheating and pressure
failure can occur within a matter of minutes. In slurry applications, where heads are
generally lower and pumps are of single-stage impeller design, the overheating takes
longer. Still, in a typical slurry application, the fluid within an isolated pump can
reach temperatures of 150–300 °C within 15–45 min depending on conditions, and
the vapor pressure of water at these temperatures may exceed that of the pump
design. Furthermore, in slurry applications, an additional mechanism for pump
isolation exists—namely, blockage of the pipeline due to the settling of solids. If a
pipeline blockage due to solids manages to seal both the suction and discharge pipe
of a pump, the potential for explosion exists.
Several strategies may be employed for reducing the risks posed by pump
blockage events:
• Monitoring of system flow rate and/or pump power. Loss of flow and/or greatly
reduced power consumption are the first and best indicators of a blocked pump.
Anytime the system flow rate goes to zero, the possibility of an isolated pump
must be considered. Slurry system operators will often attempt to clear the
blockage by increasing pump speed or taking other actions. Operators must be
aware that such measures can be applied for only a limited time before initiating
shutdown to prevent overheating of any potentially isolated pump. If the flow rate
is not monitored directly, pump power can also be used to identify a zero-flow
condition, as the pump power at the shut head is usually considerably less that the
normal operating power.
• Monitoring pump casing temperature. When strategically placed on the pump
casing, temperature probes can give an early indication of the temperature rise
that occurs in an isolated pump, allowing it to be shut down before excessive
pressure develops.
• Monitoring pump pressure and/or providing pressure relief devices. In theory,
the most direct method to avoid overpressure is to monitor the pressure, or at least
to provide a pressure relief device near the pump. It must be considered, however,
340 11 Practical Experience with Slurry Systems
that such devices may become blocked by the same solids settling event that
blocks a pipeline; therefore, they may not function properly when they are needed
most. They should, therefore, be backed up by monitoring pump flow rate, power,
and/or temperature.
• Pressure collapsing pump design. In extreme cases where pump blockages are
common, pumps may be equipped with a calibrated structural element that
collapses at a specified overpressure, releasing the pressure from the pump in a
controlled fashion before it reaches dangerous levels. These designs typically
require rebuilding of the pump after the collapsing element is activated, which can
lead to nuisance shutdowns due to “normal” overpressure events, such as a mild
water hammer. However, for certain duties, they provide a desirable safety
option.
Ultimately, an isolated pump always results in the system flow going to zero, a
condition that is usually possible to recognize on short notice. If the system flow ever
drops to zero with pumps still operating, this represents a potential hazard that must
be taken seriously and addressed quickly.
For a more in-depth treatment of this topic, with a special focus on the slurry
pump industry, see O’Conner (2006).
In designing sumps and suction piping, the main considerations are providing
consistent slurry feed, avoiding blockages, and preventing cavitation of the pump.
There are numerous considerations and few fixed solutions for all cases, so it comes
as no surprise that many problems encountered in slurry systems may be traced to
unanticipated faults in the sump or suction piping.
11.9.1 Sumps
Sump design varies with the type of slurry and the service. Significant differences
may be needed to accommodate a water-based settling slurry as compared to a
viscous non-settling slurry. Agitation may be required, and this special case is treated
in Sect. 11.9.2. Frothy slurries present additional challenges, and these are covered in
Sect. 11.10.
In any sump, it is desirable to maintain at least two meters of liquid level above
the pump centerline, as well as a minimum volume in the sump equal to at least one
minute’s discharge. Maintaining an adequate liquid level helps to prevent air
entrainment due to the development of a “bathtub” vortex, which causes pump
surging, increased wear, and shock loading on the pump shaft and bearings. In
addition, the flow into the sump should discharge below the surface, as distant as
11.9 Sumps and Suction Piping 341
possible from the pump suction pipe. Baffles or weirs, separating the inlet and outlet
regions of the tank, can be placed in the sump if air entrainment is likely to be a
problem. Cruciform baffles can be placed in the suction line to dampen the “bathtub”
vortex if present, but such flow obstructions should be avoided unless absolutely
necessary. In general, simple designs are preferred because they suffer less wear and
require the least maintenance.
The tendency of a slurry to adhere to the sump walls is an important factor.
“Sloughing off” of solids accumulated in the sump is to be avoided because it
imposes a sudden load on the pump and can upset the operation of the whole slurry
system. Tapered and rounded sides can be used to minimize solids accumulation, but
the degree of taper must be carefully chosen according to the properties of the solids,
including their natural angle of repose and their tendency to adhere to surfaces.
Improperly designed, tapered sides can lead to problems if the solids adhere to the
tapered surfaces and periodically release in large “slugs.” Alternatively, a flat-bottom
sump can be used, where the solids are allowed to build up at their natural angle of
repose. Both types of sumps are commonly used, and the choice is largely one of
preference and experience.
Figure 11.13 shows a flat-bottom sump layout, which embodies numerous fea-
tures that may contribute to good operation. The pump suction pipe passes through
the side of the sump and is entered through a downward-pointing bend with a short
bellmouth. A priming jet may be located facing into the bellmouth entrance to assist
in pump priming and to eliminate blockages on start-up. If the slurry has a high
solids concentration, or if the solids are very coarse, the priming jet may be allowed
to run continuously to prevent sporadic plugging of the inlet with resultant surging.
Water should be provided to the sump in sufficient quantity to fill the system and act
as makeup water. For in-plant systems, tank level control is sometimes used to
regulate the water flow.
Tapered-bottom sumps, as seen in Fig. 11.14, are often used to save space and
ensure the transport of settled solids to the pump inlet. The same criterion regarding
minimum volumes and level depth apply to these smaller sumps, which often require
suitable instrumentation and actively controlled feeds to prevent overflow or
draining.
Agitated sumps are used to ensure homogeneity of feed to the pumps for multi-
component slurries that segregate, or for slurries where chemical reactions may be
taking place, the latter of which can be highly viscous. For rapidly segregating
suspensions, a modification to the standard “square” mixing tank is shown in
Fig. 11.15.
Rather than using the smooth-shaped tank bottom that is popular in the chemical
industry, conical fillets have been added to prevent solids accumulating either in the
corners of the tank or directly below the agitator. Such fillets are simple to fabricate
342 11 Practical Experience with Slurry Systems
and to replace in the event of high erosion. They also have the advantage of
providing extra turbulence in the tank to assist solids suspension. The self-cleaning
baffles (seen in Fig. 11.15) may be omitted with considerable savings in power
consumption, if the suspension has a yield stress in excess of 30 Pa, provided the
tank will not be used with lower viscosity suspensions.
High viscosity, non-Newtonian suspensions with a yield stress require special
consideration. For single agitator configurations, a sheared cavern forms around the
agitator, as illustrated in Fig. 11.16a, and the boundary of this sheared area will
extend to the point where the pressure developed by the impeller balances against the
yield stress of the slurry. Using X-ray techniques, Elson et al. (1986) were able to
characterize the cavern size for various visco-plastic fluids as follows:
11.9 Sumps and Suction Piping 343
8
< D Re < 30
a
Dc = 1:36Po ρm N D2 2 ð11:3Þ
: Re ≥ 30
π2 τy
0:4Dc Dc < T
Hc = ð11:4Þ
/ N Dc T
where Re is the Reynolds number based on the agitator diameter and rotational speed
and Po is the agitator’s power number, Po = P/ρmN3D5, where P (W) is the shaft
power and N (rad/s) is the agitator’s rotational speed.
Operating tanks in this mode with reliability is difficult, and potentially danger-
ous, since no motion is observed at the surface. Operators may be tempted to
increase the mixing power in order to create visible mixing, but since mixing will
be confined to the sheared volume, overheating and boiling of the material can occur
very rapidly. By using multiple agitators, arranged at a sufficient distance such that
their flows combine, the cavern formations can be overcome, resulting in solenoidal
344 11 Practical Experience with Slurry Systems
flow in the tanks and Continuous Stirred Tank Reactor (CSTR) performance (Pullum
et al. 1994). A multiple agitator, high aspect ratio configuration that provides such
mixing, typical of mega-liter tanks found in refineries, is shown in Fig. 11.16b. The
dimensions should only be taken as a guide, as they are a function of the slurry
rheology (in this case, τy ~ 30 Pa). Such tanks are often run at various working
levels, but care must be taken to ensure that a minimum submerged depth of the top
agitator in use is maintained, as indicated. When the level falls below this, the axial
flow into the agitator becomes predominantly radial, and the top agitator no longer
pumps.
When pumping slurries, the piping between the sump and pump inlet must be
properly sized to ensure a suitable velocity for the suspension of solids entering
the pump. If the piping is too large, solids may settle, leading to adverse inlet
conditions that affect the NPSHA at the pump and possibly the wear life of the
pump impeller and suction side liner. If the piping is too small, high velocities into
the pump may cause excessive wear on the back shroud of the impeller, and pressure
11.9 Sumps and Suction Piping 345
Fig. 11.16 Mixing patterns in visco-plastic fluids: (a) cavern formation; (b) operational zones for a
multi-agitator installation showing varying successful and unsuccessful levels of fill
losses will be higher, again reducing the NPSHA. Since these sizing considerations
may vary depending on the slurry properties and flow rate, some slurry pumps have
multiple suction diameters available.
Inlet piping between the sump and pump should be kept as short as possible to
maximize the NPSHA at the pump, while allowing sufficient space for safety during
maintenance operations and stabilization of the flow before reaching the pump. In
most cases, five to ten pipe diameters are sufficient. If diffusers or reducers are used
in the suction line, they should have an included angle of no more than 15o, and if air
is present in the slurry, they should be arranged to avoid the entrapment of air at the
top of the piping.
Many slurry pumps operate without sumps, removing material directly from an open
pit or natural body of water. These can be found in the mining of sand and
aggregates, various industrial mineral deposits such as phosphate and titanium,
and of course in the many types of dredging performed for waterway maintenance,
port construction, or environmental remediation.
Open pit sump operations often have a pump located above ground level, which
draws slurry from an artificially constructed pit. Material is dumped near the pit and
drifted toward the pump suction using process water. A large screen or grizzly may
be positioned between the pump suction and the mined material in order to remove
oversize rocks and tramp, such as the fossil bones and Civil War cannonballs
346 11 Practical Experience with Slurry Systems
Fig. 11.17 Schematic of a dredge ladder pump and related variables determining NPSHA
The calculation of NPSHA was covered in Chap. 8 for the general case. However,
some additional explanation is merited for a ladder pump. Figure 11.17 shows the
ladder and defines the quantities required for NPSHA calculations. The greater the
submergence of the ladder pump, the greater the NPSHA. Thus, the ladder pump will
usually run without cavitation and will prevent cavitation in the next pump in the
line, which is normally on board the dredge. Capital and maintenance costs for the
underwater ladder pump are justified by the increased production and reduced
specific energy consumption associated with operation at higher solids concentra-
tion. Reliable handling of most solids types, at volume concentrations up to or
beyond 35%, can usually be achieved with the ladder pump configuration.
Trailing suction hopper dredges employ a vacuum type of suction draghead in
order to maximize the pick-up of fine solids. In these designs, the suction pipe often
passes through the hull at a pivot point on the side of the vessel and is raised and
lowered by winches located at the side or stern. The pumps are located low in the
hull and made large and slow running to improve their NPSHR performance. Slurries
with 40% solids by volume can be handled by this type of dredge.
Blockages in the pump suction for open pit and dredge operations can be caused
by rocks, clay balls, tree roots, vegetation, and various discarded manmade objects,
from construction materials to bicycles. The severity of the problem depends on the
size of these solids and how frequently they occur. Some sort of coarse screen or
grizzly may be useful, as noted previously, but it is often necessary to compromise
the hydraulic performance of the pump and the location of its design point to ensure
that it can pass the largest possible solids. Increasing the solids passage size of a
pump usually entails reducing the number of vanes in the impeller. It may also be
necessary to increase the passage width of the impeller, which will shift its best
efficiency point to a higher flow rate, often above the desired duty flow. Although it
is sometimes possible to counteract this effect with a specially designed casing, it is
not uncommon to find large pumps operating well below their best efficiency flow
rate with low efficiency and poor wear life in order to eliminate the risk of blockage.
348 11 Practical Experience with Slurry Systems
Fig. 11.18 Root cutter installation (looking into the pump suction)
When roots and vegetation are the main concern, some operators use a so-called
“root cutter,” intended to keep impellers from clogging. This device consists of a bar
bolted inside the suction pipe, running parallel to the pipe axis, and extending into
the eye of the impeller, close to the leading edge of the impeller vanes. Figure 11.18
shows an example of a root cutter looking through the suction pipe into the pump
suction. Root cutters place significant loads of the pump impeller and rotating
assembly and can easily cause impeller breakage or bearing failure. It is therefore
no surprise that they are generally not sanctioned by the equipment manufacturer.
Nonetheless, they are considered essential in some applications by the operators.
Tests show that the associated reductions in head and efficiency are generally
negligible, but the effect on suction performance can be important, generally
resulting in a 20% increase in the pump’s NPSHR.
Jet pumps may be used at the entrance to a suction piping system, or even at the
suction of a centrifugal pump, to prevent blockages and improve solids feeding. A jet
pump adds to the NPSHA and total pump head; however, such devices are usually
less than 50% efficient when considering the effect of the supply water pump and its
drivers. The water they add also reduces the solids concentration and has an adverse
effect on specific energy consumption. Nevertheless, jet pumps do eliminate some
wearing parts, and they have a role to play in handling difficult slurries. They are
most likely to be useful where dilution can be tolerated, operating efficiency is
unimportant, and the system head is low.
11.10 Pumping Frothy Mixtures 349
Despite efforts to eliminate air entrainment, frothy mixtures are often encountered in
slurry pumping. Common causes include excessive turbulence of the flow entering
the pump, or a sump level that is too low. Other sources are chemical reactions, the
release of gases contained in the slurry solids, and the intentional introduction of air
during mineral separation by flotation processes. All of these problems may be
compounded in agitated sumps.
When gas is present, the density of the slurry mixture is reduced, and the pressure
produced by the centrifugal pump is reduced in direct proportion. This effectively
reduces the head produced by the pump, if head is calculated according to the density
of the de-aerated mixture, even if the head of the actual mixture (including gas)
remains unchanged.
When the volume of gas exceeds some critical percentage of the mixture, excess
gas will begin to accumulate within the pump suction and impeller passages,
resulting in a blockage that significantly reduces the head, efficiency, and NPSHR
performance of the pump. Gas bubbles in liquids will always be driven to the area of
lowest pressure by the forces of buoyancy acting upon them; in a centrifugal pump,
this low-pressure area is represented by the suction inlet of the impeller. This
accumulation begins when the buoyant forces directing the gas bubbles toward the
pump suction cannot be overcome by the opposing drag forces created by the flow
through the pump. The onset of the blockage is often sudden rather than gradual once
the tipping point is reached between these opposing forces. The resulting reduction
in pump head can be large—as much as 30–40%—and the impact of such a sudden
change on the operation of the slurry system can be significant. Furthermore, since
the pump’s NPSHR is increased by the blockage, attempts to recover the lost head by
increasing pump speed may not be successful; since the greatly increased NPSHR
which results from the combination of the blockage and the increased pump speed,
will often exceed the NPSHA of the system. The system then cannot be driven by the
available pumping capacity and will appear “locked.” This condition is often
referred to as an “air lock” or “air binding.”
Once the introduction of gas is reduced or eliminated, the blockage may clear;
however, in some cases (in particular with viscous mixtures), it may be necessary to
shut the pump down and allow the excess gas to escape before the pump can be
restarted. In these cases, it is important to orient the pump discharge position and
slope of the discharge piping so that gas can escape some distance away from the
pump casing. Otherwise, it may simply be sucked back into the impeller, resuming
the air-locked condition once the pump restarts.
The volume percentage of gas at which air lock occurs varies widely according to
the pump design and the slurry mixture’s physical properties. In clear water appli-
cations, the onset may occur with as little as 3% gas. In froth process applications,
where chemicals are added to modify the surface tension of the fluid and fine solids
are present, gas volume percentages of 15% or more may be handled without air
lock. Unfortunately, these same additions also encourage the development of very
350 11 Practical Experience with Slurry Systems
Fig. 11.19 Examples showing the onset of air lock for an unvented froth pump design
frothy mixtures that contain high volumes of gas and break down only slowly over
time. As a result, many froth-handling pumps in process applications operate in a
continual state of air lock. Figure 11.19 shows some examples of pump performance
due to air lock, the nature of its sudden onset, and the effects of flow rate and surface
tension modifying chemicals.
Whenever possible, the first and most effective strategy for improving pump
performance in such conditions is to reduce the amount of gas reaching the pump. If
the gas is being introduced by poor sump design, insufficient sump level, or some
other fault in the suction piping design, the issue should be corrected. If the mixture
is naturally frothy, or if it is made frothy by the process itself, steps may be taken to
reduce the volume of gas in the froth before it reaches the pump. These may include
allowing the mixture to run some distance in shallow, open launders, possibly
accompanied by the application of a fine spray of process liquid over the top of
the froth. A riser may also be installed in the suction piping, just before the pump, to
capture the larger gas bubbles and vent them to the atmosphere, or—better yet—to a
continuously evacuated chamber held at some low pressure (Herbich 1975). To be
effective, the intersection of such risers with the suction piping must be large, ideally
spanning the entire width of the pipe and a length of at least two to three pipe
diameters, as shown in Fig. 11.20.
Many specialized sump designs have been devised to minimize air entrainment
when the slurry is inherently frothy. Common strategies involve baffles to create a
tortuous path for the slurry, and the placement of vents in strategic areas to allow
accumulated air the opportunity to escape. The variety of these designs is too broad
to cover here, but a key success factor in all such designs is the areas of reduced flow
velocity where entrained gas can rise to the surface and escape. If flow velocities
11.10 Pumping Frothy Mixtures 351
within the sump remain high, gasses will remain entrained, no matter how clever the
design. Conical sump designs with a tangential entrance flow to encourage the
formation of a vortex within the sump are also used, on the premise that the
lower-density air will be centrifuged to the air core of the vortex and escape from
the slurry. However, these also require a sufficient size and residence time to be
effective. For more viscous slurries, or ones where the interface forces are high,
shear can be induced to assist in the separation of the air (see Sect. 2.7), either
through high shear flows or through mechanically induced shear (Furlan et al. 2014).
Figure 11.21 shows a very simple shear-inducing device that is designed to remove
air from such slurries.
Pump designs may also be modified to accommodate a larger volume of gas
before the onset of air lock. In many cases, open-shrouded impeller designs (i.e., an
impeller without a suction side shroud) are used in froth applications. The extra
turbulence created by these designs appears to delay the onset of air lock to some
higher volume of gas—for example, from 5% to 8% in clear water applications.
Extension of the vanes into the suction piping can produce additional turbulence and
further delay the onset of air lock in viscous applications.
An inducer can also assist in froth-pumping applications. The inducer is essen-
tially a propeller, or axial pump, attached to and turning with the impeller, and
extending into the suction pipe just ahead of the pump (see Fig. 11.22). Inducers are
commonly used to boost suction pressure and NPSHR performance in high-head
centrifugal pump applications in the energy and water industries. Inducers have two
disadvantages, however, which make them ill-suited for slurry applications. First,
they are only effective within the narrow range of flow rate for which they are
designed. Outside this range, they produce a pressure loss instead of providing a
boost. In many slurry applications, the operating flow varies widely, and maintaining
the correct flow rate for effective inducer operation may be difficult. The second
disadvantage is that inducers require a very close clearance between their outer
diameter and the inner diameter of the suction pipe, and this clearance is difficult to
352 11 Practical Experience with Slurry Systems
maintain in slurry applications where solids are present. Clearances greater than
1 mm can result in pressure loss, rather than boost, at the suction inlet.
Vented pump designs have been developed to continuously remove the accumu-
lation of gas from the pump suction, as shown in Fig. 11.23. While these designs do
not prevent the onset of air lock, they can control the pump head loss to a more
manageable level of 5–15%. They can only be applied within a specific range of
suction pressure, which must be above atmospheric pressure for the venting system
to work (i.e., no suction lift) but also low enough that fluid is not pushed out of the
venting system. As a result, they are generally limited to single-stage (or first-stage)
applications. However, many froth applications in mineral processing meet these
conditions and can be improved with the application of vented designs. For a more
detailed treatment of vented pump designs see Visintainer and Whitlock (2012).
A pump’s operating condition may also affect the onset of air lock. In particular,
operation at a higher suction inlet velocity may delay air lock by exerting higher drag
forces on the gas bubbles to counteract the buoyant forces driving them toward the
suction. Unfortunately, this approach also results in a smaller suction inlet area,
which runs counter to the practice of oversizing pumps according to the froth factor
strategy outlined below.
In some cases, air lock can be hard to distinguish from cavitation. One helpful test
for distinguishing between the two is to increase the system resistance downstream
of the pump by a valve or other means and look for either an increase in pump
discharge pressure or a decrease in noise. Since reducing flow will reduce NPSHR
and increase NPSHA, cavitation will often disappear under these conditions, and the
pump performance will exhibit a noticeable recovery, and possibly also a reduction
in pump noise and vibration. If, however, the pump is experiencing air lock, a
reduction in flow rate will often make matters worse.
354 11 Practical Experience with Slurry Systems
Another consideration for the design of a slurry system with gas present is the fact
that the volume of gas will change with system pressure. In particular, the gas
volume at the pump inlet will generally be less than seen at the top of the sump,
due to the static pressure of the fluid above it. Therefore, a deeper sump is always an
advantage. Also, the gas volume will change significantly between the suction and
discharge of the pump, such that the gas volume in the discharge piping may be less
than half of that in the suction piping. The actual values can be calculated according
to the well-known gas laws. While such corrections may be neglected in short
systems with only one pump, they should not be ignored for multi-pump systems.
One positive benefit of this reduction in gas volume is that downstream pumps in
multi-pump systems may be less likely to experience air lock and may therefore not
require specialized designs.
Regardless of the above strategies, many applications remain where pumps must
run under continuous air-lock conditions. This may be because the actions for
preventing it are not feasible, or because they are considered too expensive—or it
may simply be due the tenacious nature of the froth itself. In these cases, the pump
selection must be made such that the system will continue to operate as desired, even
when the pump is in an air-locked condition. The most common method to achieve
this end is to apply a froth factor to the flow rate when making the pump selection.
This froth factor is simply a multiplier applied to the process flow rate when selecting
the pump, so that the pump selected is a larger one that can deliver the required flow
rate in the air-locked condition. The exact value must usually be determined by
experience with similar slurries, although some rough guidelines exist. Common
froth factors are in the range of 1.5–2.5, but factors as high as 8 are used in the most
difficult applications. A detailed procedure for this selection process, along with
guidance regarding froth factor values, may be found in the Hydraulic Institute
Slurry Pump Standard ANSI/HI 12.1–12.6 (2021).
ρm gQH
Power = ð11:5Þ
η
where ρm is the slurry density, Q is the flow rate, H is the total developed head in
height of slurry, and η is the pump efficiency expressed as a fraction.1 For many
water and process pumps, H falls off sharply as Q increases, so that the pump power
1
Eq. (11.5) applies in S.I. or any other consistent set of units. If mixed units are used, as is
customary in the U.S., an additional numerical coefficient is required.
11.11 Slurry Pump Drive Trains 355
goes through a maximum. However, slurry pumps typically have flatter head-flow
characteristics, so that the power demand increases monotonically with Q. The
power demand also increases with slurry density; therefore, the maximum power
often occurs at the maximum required flow rate and/or slurry density. The value of η
must include the effect of solids on pump efficiency, discussed in Chap. 9. Further-
more, the motor rating must allow for losses in the motor and drive.
For a fixed-speed, direct drive to be used, the motor speed must be suitable, the
pump and motor must be closely matched, and operation must be steady. These
conditions are rarely met in slurry systems. Therefore, drive trains for slurry pumps
often consist of a motor operating at a relatively high speed with a speed reducer, or
some type of variable-speed drive. In many cases, variable-speed drives and speed
reducers are used together. Where fixed-speed pumps are used, it is often necessary
to add “makeup water” to the sump in variable quantities to adjust for variations in
the feed conditions.
The pump and drive must be able to respond to required changes in operating
conditions. Some implications for system design and pump selection are considered
in more detail in Chaps. 10 and 14. Where the solids throughput or size varies
widely, the power requirements will also vary widely and may change suddenly.
This is typical of many mining and dredging operations, where substantial power
changes can occur on a minute-by-minute basis. When there is a high degree of
variability or uncertainty in the input conditions, the pump speed must also be
variable. Characteristic curves for a variable-speed pumping system are shown in
Fig. 11.24.
Four main types of drive are in common use for slurry pumps: fixed-speed electric
motors, hydraulic motors, variable-speed electric motors, and piston engines. With
fixed-speed motors, pump shaft speed can be varied using gears or V-belts. V-belts
are relatively inexpensive and simple to maintain, and changes in speed can be
executed with relatively little effort. Single-reduction gearboxes are often used on
larger units, and the designer must then remember that the pump will rotate in the
opposite direction to the drive motor. For the rough and variable operating condi-
tions often encountered in mining and dredging, an extra service factor on their
power rating is recommended, and values of 1.5–2.0 times the delivered power may
be appropriate.
The hydraulic motor offers the advantage of variable speed. It is usually relatively
inexpensive and easy to maintain but is limited in power. Its major disadvantages are
its limited speed range and low drive efficiency. This drive system is normally
confined to small dredges.
Solid-state variable speed drives have become increasingly common, replacing
both fixed-speed installations and older inefficient designs involving slip-ring or
DC-shunt motors. These drives are often implemented using a 4-pole motor with a
gearbox or belt drive, although direct coupling of lower speed (higher pole number)
motors to slurry pumps is becoming more common. This gives the advantage of
eliminating the gear reducer, but can result in constrained flexibility for future speed
changes which may have been accommodated by changing gears. Solid-state drives
may be either AC or DC. The AC type is more common because it uses a near-
standard motor. The DC type has traditionally had the advantage in that its operation
can provide constant torque to lower speeds, but this feature is of limited value for a
centrifugal pump, where the torque required at low speed is usually also low. The
efficiencies and costs for both types continue to improve, justifying their use in more
applications.
Dredge pumps are often driven by a diesel engine. Some manufacturers provide
these engines to cover a wide range of power ratings, and the adaptation to a dredge
pump is usually through a gear reducer. In choosing an engine of this type, some
attention must be given to its speed-torque characteristics. If properly configured, the
diesel engine can function as a variable-speed, constant-torque drive over a very
wide range of conditions. When a wide operating range is required, turbocharged
four-stroke diesels appear to have an advantage over the two-stroke types. Some
engines can operate at full torque between 80% and 100% of full-rated speed. This
20% speed range is usually equivalent to a variation in pump head of about 35%,
allowing operations over a wide range of conditions without the need to change
impeller size. This arrangement is able to provide virtually the full available power
over a broad range of conditions. Correct controls on the engine are important, and
the engine manufacturer should be fully briefed as to the requirements, so that
correct governor settings can be established.
Drives for ladder pumps on dredges are sometimes arranged with the motor
mounted above the water surface and a long drive shaft passing down the ladder.
Modern arrangements tend to use submerged electric motors, sometimes variable-
speed DC, with a characteristic of the type shown in Fig. 11.25.
11.11 Slurry Pump Drive Trains 357
Fig. 11.26 Typical Florida phosphate matrix transport pump. (Courtesy GIW Industries, Inc.)
References
ANSI/HI 12.1–12.6, American National Standard for rotodynamic centrifugal slurry pumps
(Hydraulic Institute, Parsippany, 2021)
A. Kodura, K. Weinerowska-Bords, W. Artichowicz, M. Kubrak, P. Stefanek, In situ verification of
numerical model of water hammer in slurries. ASME J Fluids Eng 141(8), 081115 (2019)
B.P. O’Conner, Centrifugal pump explosions. Presented at the Conference of the South African
Institute of Mining and Metallurgy, Johannesburg, 2006
B.B. Sharp, Water hammer: Problems and solutions (Edward Arnold, London, 1981)
C. Cristoffanini, M. Karkare, M. Aceituno, Transient simulation of long-distance tailings and
concentrate pipelines for operator training (Presented at the SME Annual Mtg. & Exhibit,
Salt Lake City, 2014)
D.J. Wood, T. Kao, Unsteady flow of solid-liquid suspensions. J Eng Mech ASCE 92, 117–134
(1966)
E.B. Wylie, V.L. Streeter, Fluid transients (McGraw-Hill, New York, 1978)
I.J. Brewster, Tailings disposal pipeline for bougainville copper limited. Aust Civ Eng Trans CE35,
325–334 (1993)
I. Cave, Slurry turbines for energy recovery. Presented at the 7th International Conference on the
Hydraulic Transport of Solids in Pipes, Sendai, 1980
J.M. Furlan, R.J. Visintainer, A. Sellgren, Centrifugal pump performance for highly non-Newtonian
clays and tailing slurries. Presented at the 19th International Conference on Hydrotransport,
Golden, 2014
J.B. Herbich, Coastal and deep ocean dredging (Gulf Publishing Company, Houston, 1975)
J. Parmakian, Waterhammer analysis (Prentice Hall, New York, 1955)
J. Rusconi, A. Lakhouaia, M. Kopuz, The design and engineering of the 187 km Khouribga to Jorf
Lasfar phosphate slurry pipeline. Procedia Eng 138, 142–150 (2016)
L. Pullum, M.C. Welsh, K. Baillie, P. Kam, A Flow visualisation study of mixing solid/liquid
slurries by mechanical agitation (Presented at the ASME 5th Intl. Symposium on Solid-Liquid
Flows, Los Angeles, 1994)
References 359
M.R. Carstens, T.W. Hagler, Water hammer resulting from cavitating pumps. J Hydr Div ASCE 90,
161–184 (1964)
M.H. Chaudhry, Applied hydraulic transients, 2nd edn. (Van Nostrand Reinhold, New York, 1987)
M.W. Chudacek, Relationships between solids suspension criteria, mechanism of suspension, tank
geometry, and scale-up parameters in stirred tanks. Ind & Eng Chem Fundm 25, 391–401
(1986)
M. Wilkens, C. Gilchrist, M. Fehrsen, R. Cooke, Boulby mine backfill system: Design, commis-
sioning and operation. Presented at the 16th International Conference on the Hydraulic Trans-
port of Solids, Santiago, 2004
R.G. Betinol, H.E. Jaime, Startup of dual concentrate pipeline for Minera Escondida Limitada.
Presented at the 16th International Conference on the Hydraulic Transport of Solids,
Santiago, 2004
R.H. Derammelaere, G. Shou, Altamina’s copper and zinc concentrate pipeline incorporates
advanced technologies. Presented at the 15th International Conference on the Hydraulic Trans-
port of Solids, Banff, 2002
R.J. Visintainer, L. Whitlock, Development and testing of a more effective froth handling pump.
Presented at the 44th Canadian Mineral Processors Conference, Ottawa, Ontario, 2012
T.P. Elson, D.J. Cheesman, A.W. Nienow, X-ray studies of cavern sizes and mixing performance
with fluids possessing a yield stress. Chem. Eng. Sci. 41, 2555–2562 (1986)
W. Han, Z. Dong, H. Chai, Water hammer in pipelines with hyperconcentrated slurry flows carrying
solid particles. Sci China, Series E 41(4), 337–347 (1998)
Chapter 12
Testing and Instrumentation
12.1 Introduction
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 361
R. Visintainer et al., Slurry Transport Using Centrifugal Pumps,
https://doi.org/10.1007/978-3-031-25440-6_12
362 12 Testing and Instrumentation
specific test programs may include carrier fluid viscosity and density, onset of
stationary deposition, and local wear rates. More specialized measurements may
include detailed concentration, particle size and velocity distributions within the
flow, chemical properties (such as pH and conductivity), and the slurry’s behavior
under special conditions, such as a prolonged shutdown. For more information on
the instruments most commonly used for slurry testing, see Sect. 12.5.
How Will the Slurry Be Obtained, Prepared, and Shipped to the Test Location?
Slurry components for testing may be taken directly from existing deposits or
production streams, or may be specially made in pilot plants. It is always important
to consider whether the components accurately represent the end use the testing is
intended to simulate.
If shipped already mixed, slurries with broad particle size distributions have a
tendency to settle into very compact layers. Remixing may be difficult without
damaging the properties of the solids. If solids and fluids are shipped separately,
special provisions may be needed for combining them without entraining air or
producing clumps of aggregated solids. If flocculants or other chemicals that alter the
slurry properties are to be tested, careful consideration must be made about whether
to ship the slurry pre-mixed, or to add the chemicals at the time of testing. In
addition, if the slurry contains any hazardous components, arrangements must be
made for proper handling, containment, and disposal.
A great deal of experienced judgment is usually required to assess the above
points and develop a cost-effective test program that will deliver useful information
to the slurry system designers and operators. As such, it is highly recommended that
personnel with prior experience in slurry testing be engaged for all but the simplest
test programs.
Once a test plan has been developed, a range of testing strategies is available for
collecting the required data. These strategies may be used individually or in concert,
depending on the purpose of the test program and the importance of the results.
Important features and considerations for the three most common approaches—
bench, on-site, and laboratory pipe-loop testing—are described below.
Bench Tests
Some of the properties to be measured during a slurry test either cannot be obtained
in real time, or may not change much during testing. These kinds of properties are
typically determined using bench-top devices from slurry samples gathered before,
during, or after testing. These often include the density of the solids, the particle size
distributions, and the density and viscosity of the carrier fluid.
12.2 Pipeline Testing 365
In the special case of homogeneous and non-settling slurries, data sufficient for
pipeline design can often be obtained using bench viscometers of the type discussed
in Sect. 12.4, without the necessity of a full series of pipeline tests. For more
information on these instruments, see Sect. 12.4.1.
Bench scale tests can be conducted on-site or in the laboratory, but on-site testing
has the advantage of the ready availability of the actual materials in the desired
conditions.
On-Site Testing
On-site testing is often limited by production schedules and plant layout; however,
attempts should be made to obtain data at the widest range of conditions, rather than
restricting data collection to the particular conditions of practical interest. This is
because many slurry properties are best understood in the context of their behavior
under changing conditions, and in any case, all processes see variability in practice
that designers and operators must anticipate. It is important that the production
control room is made aware of the tests so that any events can be logged.
Much of the instrumentation required may already be installed for production
control purposes. While convenient, the accuracy of these instruments should be
considered and, if possible, new calibrations made before the start of testing, since
the accuracy requirements acceptable for process control are often less than those
required for a successful test program. Some additional instrumentation may be
needed to monitor key variables (in particular, pressure measurements throughout
the system) that are often not needed for process control. Access to the various
instrument time histories is usually available after the tests, but whenever possible,
live data should be obtained, either by reading instruments directly or from control
room screens. These data can be plotted directly, providing the test team with
immediate insight into the flows and some assurance that valid data are being
collected. This is particularly important for coarse settling suspensions at flow
rates near deposition, where the danger of blocking a line and disrupting production
is high. Since any site work has its own restrictions and peculiarities, further advice
cannot be given, but examination of some successful procedures described below for
laboratory testing can provide an idealized plan of action for on-site work.
Laboratory Pipe-Loop Testing
Unlike in pump performance testing, there are no industry standards to guide the
researcher when trying to establish slurry transport characteristics. In addition, test
equipment will vary from one establishment to another, depending on the goals and
constraints of each facility. These facts highlight the importance of the test plan
development phase outlined previously. However, certain elements are commonly
present, as shown in Fig. 12.1.
The slurry holding tank often doubles as a sump for the slurry pump. When sizing
the tank, test planners should consider both slurry degradation and the importance of
collecting realistic data in the range of solids stratification and/or deposition. A larger
sump will reduce the frequency of circulation for individual particles and provide an
inventory of coarse solids for the pipe loop to draw upon, as these solids will form
heavily stratified or stationary beds within the pipe. However, the larger sump will
366 12 Testing and Instrumentation
sample cutter
optional density
agitator meter
temperature
probe optional optical flow meter
window
also require a larger slurry sample. In practice, tank-to-pipe-loop volume ratios will
vary from two to ten, depending on the importance of these factors. For settling
slurries, the turbulence provided by the returning flow is usually sufficient to
maintain the solids in suspension in the tank, although for coarse solids, this may
be less than optimum. The returning pipe should enter near the top of the tank (below
the waterline, to avoid air entrainment), and the feed pipe should exit from the
bottom to maximize residence time within the sump and ensure that the coarse solids
are not trapped in the sump. For viscous non-settling slurries or pseudo-
homogeneous slurries, some means of removing air from the system (inevitably
entrained during loading) should be provided. A suitable agitator or special tank
design may also be needed to suspend the particles and ensure that uniform feed
conditions are maintained.
The pump should have sufficient power to deliver the highest flow rates required
at the highest expected concentration of solids, and it should be equipped with a
variable speed drive to avoid the necessity of a control valve to vary the flow rate
during testing. Ideally, the pump should have a mechanical seal to prevent gland
water from entering the test loop and diluting the slurry, unless this additional water
can be drawn off elsewhere in the system without altering the slurry properties.
Dilution control is especially important when testing non-settling, non-Newtonian
materials. A flow meter suitable for slurries (see Sect. 12.5.1) should be installed in a
vertical section at least ten diameters downstream of any bend.
Piping should be flanged for easy access, and made of materials that can with-
stand the expected operating pressures and meet any chemical requirements. The
pipe size in laboratory testing is often smaller than the end use for which data are
being collected, so the piping should be selected according to the considerations of
scalability covered in previous chapters This is especially important for testing
complex, non-Newtonian mixtures, as described in Sect. 6.3.1. In general, the
recommended pipe size for homogenous slurries is at least 75 mm, and for complex
slurries at least 100 mm. Larger sizes may be needed in some cases, depending on
the specifics of the slurry (such as the largest particle size) and the desired end use for
the data.
12.2 Pipeline Testing 367
It is necessary to obtain accurate, clear water (or clear carrier fluid) data as a
baseline, so it is usually necessary to scour the system with a high-velocity sand
suspension before testing to remove any corrosion or deposits and bring the inside
surface of the pipe into a “steady state” of smoothness (i.e., one that will not change
further during testing). If the test facility is used to examine non-Newtonian slurries,
it is advisable to construct the entire system from a combination of plastics (e.g.,
ABS or HDPE) and 316 stainless steel (or equivalent) to avoid contamination and
degradation of the slurry rheological properties.
For best accuracy in the laboratory, “wet” pressure tappings should be used,
where a fluid-filled line connects the pressure transducer to the system through a
small hole or “tap” in the pipe wall. (See Sect. 12.5.2).
Pressure transducers should be compatible with water and have very small swept
volumes; this helps prevent solids from being induced into the pressure tappings
themselves. For general purpose work, transducers with frequency responses of
order 3–5 Hz, typical of industrial transducers, are adequate. Differential pressure
transducers, “daisy chained” along the test section, as shown in Fig. 12.1, are
preferred, as they provide the most information and are the most accurate, measuring
the relatively small frictional component of the transport gradient while being
insensitive to fluctuations in the overall static pressure. However, at least one static
and preferably several static transducers, deployed along the test section, can be used
at the expense of some fidelity. All pressure transducers, whether static or differen-
tial, should be connected to full- or partial-pressure rings. These are ideally
supplemented with settling pots to collect any solids that find their way into the
measuring system, thus preventing pressure line blockage and protecting the trans-
ducers from solids migration (see Fig. 12.9). Tappings should also not be located on
the top of the pipe, but off center. This reduces the chance of air entrainment into the
tapping lines, which then increases the likelihood of solids entrainment as the air
volume is compressed, as well as introducing an error into the measurement when
using inclined differential pressure transducers.
An inline density meter is a useful addition, especially where the slurry density is
unknown, or may vary. If mounted in a horizontal line, the effects of solids
stratification and deposition must be considered, as these instruments usually mea-
sure the in-situ concentration (Cvi), which may include stationary deposits. Where
possible, vertical installation in established flow (i.e., at least ten diameters down-
stream of any bend or fitting, and preferably more if the slurries are rapidly settling or
non-Newtonian), will reduce particle segregation and provide improved accuracy,
especially for instruments that do not measure the entire cross-section. Vertical
installation will also usually reduce the difference between the in-situ and delivered
concentration. Where the most accurate measurement of delivered concentration
(Cvd) is desired, an inverted U-tube is recommended (see Sect. 12.5.3). An alterna-
tive method is to momentarily divert the entire flow into a vessel equipped with load
cells and volume detection. The collected sample can then be returned to the slurry
tank once measurements have been made. The volumes of such vessels need to be of
sufficient size to obtain an accurate determination of Cvd, but not so large that they
reduce the head in the slurry tank and change the pump’s performance. This is less of
368 12 Testing and Instrumentation
a problem with modern test loops, where control of the pump is provided by suitable
control algorithms that maintain constant flow conditions.
Before the slurry is returned to the slurry tank, provision should be made to
sample the slurry stream for particle size distribution and other bench-top measure-
ments. This is best achieved with a well-designed sample cutter that passes through
the entire flow, typically by sweeping a slotted collecting vessel across the entire
pipe discharge at constant velocity.
Where possible, installation of an optical window can provide great insights into
the flows, especially where transparent analog suspensions are used. Boundary
effects mean that the scouring of the windows is often not as severe as expected,
and acrylic windows can maintain sufficiently transparent surfaces for a considerable
time. However, care must be taken not to stress the windows, especially if they are
glass, and there must be suitable screening to comply with any OSHA requirements.
Where possible, the optical window may be encased in a transparent rectangular
water jacket to minimize the distortion of viewing through a cylindrical surface.
All instrumentation should be logged using a suitably equipped computer, with a
logging frequency of order 10–100 Hz for general use. Suitably averaged readings
should be displayed, preferably graphically in real time, along with the standard
deviation of the readings, to provide an indication of flow regime changes.
As test points should be taken under steady flow conditions, a closed loop is
desirable for laboratory testing. Piping arrangements for the loop depend on the type
and configuration of the pumps to be tested. Both pressurized loops and open loops
are in common use. A representative pressurized system is shown schematically in
Fig. 12.2 and illustrates some features of this type of system. The advantage of the
pressurized arrangement is that the pump suction inlet pressure can be more easily
controlled. Changes in the system pressure do not affect the total developed head of
the pump, but lowering the pressure in the system lowers the NPSHA, which
facilitates suction performance testing.
The tank is necessary to trap any air in the system. Pump performance can be very
sensitive to small amounts of air. As little as 2% by volume can lead to accumulation
of air in the suction inlet of the pump, also known as “air-lock” or “air-binding,” (see
Sect. 11.10). This air-lock acts as a blockage and greatly reduces the measured head.
Even smaller amounts of air can have a significant impact on pump suction perfor-
mance. If the test must be carried out by drawing water from an open tank, care must
be taken that the water does not become unduly aerated on returning to this tank.
With an arrangement of this type, NPSHR testing necessitates lowering the level in
the tank, which may limit the range of achievable NPSHA values.
Since the suction piping is often under negative gauge pressure during NPSHR
tests, special care must be taken to ensure that all piping joints, especially between
the tank and the pump suction, are air-tight. Joints which may not leak water under
positive pressure may still suck in air under negative pressure. Small leaks of air into
the suction piping can easily go undetected, resulting in a false determination of the
NPSHR.
370 12 Testing and Instrumentation
The piping diameter at the pump should provide a close match to the inlet and
outlet diameters of the pump. At least ten diameters of straight piping are desirable
on the suction side. Bends or tapers directly on the discharge or suction flanges of the
pump are not permissible, unless they form part of the pump design, and their effect
is considered part of the pump performance.
Variation in system resistance is best achieved by a suitable valve downstream of
the pump discharge, usually placed close to the re-entry point into the tank. This
location keeps the pressure within the tank to a minimum, simplifying its design. It
also allows any air which might enter the system through the valve flanges or valve
stem seal, due to low pressures around the valve itself, the opportunity to collect in,
and be vented from, the tank. At high differential heads, throttling valves may
experience considerable cavitation. Air bleed valves are necessary on the tank and
at several locations around the loop for use in priming and after start-up for system
deaeration.
Measurements taken during a typical H-Q-η test to determine pump hydraulic
performance include the system flow rate, the pressures on both sides of the pump,
the differential pressure across the pump, the fluid temperature, slurry density, the
pump shaft speed, and the torque or power supplied to the pump or pump driver.
Flow measurement may be carried out in several ways. The square-edged orifice
is accepted as a standard technique for flow measurement and is commonly used for
slurry pump water performance testing. However, it is not suitable if the pump is
being tested on slurry, due to the accelerated wear that the orifice would experience,
and other means such as a magnetic or sonar flow meter must be used.
Pressures are measured using both static and differential pressure transducers. A
differential measurement across the pump will be the most accurate for the head
calculation. However, stand-alone pressure measurements such as the static suction
pressure are required for the calculation of the NPSHA and the static discharge
pressure to ensure that the maximum allowable working pressures (MAWP) of the
pump and test system are not exceeded. For best accuracy, “wet” pressure tappings
are used, where a fluid-filled line connects the pressure transducer to the system
through a small hole or “tap” in the pipe wall, ideally with a “ring tap” arrangement,
consisting of four taps equally spaced around the pipe and connected by a ring of
tubing (see Sect. 12.5.2). For pump head measurements, taps should be one to two
pipe diameters from the suction and discharge flanges of the pump, in accordance
with the relevant test standard. The transducers used in pressure measurement should
be selected and arranged so that an accuracy of ±1% or better is obtained. The lines
connecting them to the system should be kept full of water, with provisions for
flushing to remove any air or solids from the lines prior to testing. For slurry testing,
it is advisable to provide settling pots close to the piping to collect any particles that
enter the pressure measurement lines, since they can introduce error by changing the
effective density of the fluid in the lines, or by blocking them altogether (see
Fig. 12.9).
The temperature of the fluid being pumped must be monitored during testing to
allow for corrections to the density, vapor pressure, and in some cases, viscosity of
the fluid. The water in the test loop should, if possible, be kept below 40 °C, since the
12.3 Pump Performance Testing 371
For more information on instrumentation, see Sect. 12.5. During data collection,
readings are recorded at 3–15 flow rates, depending on the requirements of the test
and the range of flow rate being measured, but at least eight points should be
collected when determining the full characteristic curve. The measurements at
each flow rate should be averaged from multiple readings taken after steady-state
operation has been achieved. With computer data acquisition systems, high-
frequency readings are generally averaged over 15–30 s. Total developed head,
input power, and pump efficiency are then calculated and plotted against flow rate.
Ideally, pump speed is kept constant throughout the test, but this is not always
possible due to driver limitations. Whenever necessary, values of Q, H, and P can be
scaled to a single speed using the affinity laws (Eqs. 8.6, 8.7 and 8.8).
For suction performance testing, the most satisfactory equipment is a closed-loop
system with provision for de-aeration and suction pressure control. A series of tests
may be performed, since each NPSHR test provides data at only one flow rate, and at
least five tests are recommended when determining the full characteristic curve.
Each test begins with the suction pressure at a high value, typically about one bar
above atmospheric, or three times the expected NPSHR (whichever is greater), in
order to provide an NPSHA well above any possibility of cavitation. The suction
pressure is then reduced in increments, while the pump speed and flow rate are
maintained constant. Data are collected at each change in suction pressure until
cavitation occurs. NPSHA is calculated from the measured pressure and velocity in
the pump suction inlet pipe using Eq. (8.16), and the required NPSHR for most
industrial pumps (including slurry pumps) is defined as the NPSHA corresponding to
a 3% loss of measured head, as described in Sect. 8.3. During NPSHR testing, gasses
dissolved in the liquid may come out of solution when the pressure is reduced,
effectively increasing the volumetric flow rate at the suction inlet and resulting in an
artificial increase in the determined NPSHR. To avoid this complication, the liquid in
the loop can be degassed by running a preliminary NPSHR test, heating the liquid
and allowing it to re-cool before testing, or allowing the full loop to sit static for a
day or so.
When performance evaluation is to be made using slurries, then similar questions
must be asked and remedies employed as those outlined in Sect. 12.2.1.
T T T N
N
T
N
(d)
Dp
Q
(e)
Fig. 12.3 Appropriate rheometer geometries for analyzing slurries. Measured variables are
N (rotational speed), T (torque), Δp (pressure drop), and Q (flow rate). The geometries are rotating
cup & bob (a) Searle, (b) Couette, (c) vane, (d) parallel plate, and (e) capillary tube
Many methods have been devised to measure the viscosity of a fluid and/or to
approximate the infinite flat plates used in defining Newtonian viscosity. Many of
these devices are useful for comparative measurements in the context of a process or
quality control, but only a few types can be used if scientific data are to be gathered
for use in pump and pipeline calculations. This restriction is compounded for
slurries, where instrument dimensions must be sufficiently large to accommodate
the particles in suspension. Suitable rheometer geometries for slurries are shown in
Fig. 12.3.
The cup & bob geometries described in Fig. 12.3a, b approximate the infinite flat
plates illustrated in Fig. 2.1 by using a small annular gap between an inner bob and
outer cup, one of which rotates while the other is stationary. Rotation of the bob is
the most convenient setup, but cup rotation is more suitable for lower-viscosity
materials, as Couette instabilities are suppressed in this mode. In both cases, if the
gap is small enough, the shear stress across the gap may be assumed constant and the
velocity profile linear. For homogeneous fluids, the size of this gap is restricted only
by manufacturing constraints. For slurries, the gap must be larger than at least three
times the diameter of the largest particle, and a value of ten is commonly used.
374 12 Testing and Instrumentation
where cv and cvmax are the solids volume concentration and maximum solids
concentration, respectively.
The vane type geometry shown in Fig. 12.3c is used to measure the yield stress of
a slurry and should only be used in an “infinite sea” (i.e., where the vane dimensions
are small compared to the slurry container). Suitable dimensions for the vane and
cup are H/D < 3.5, DT/D > 2, Z1/D > 1, and Z2/D > 0.5 where H and D are the vane
height and diameter, DT is the cup diameter and Z1 and Z2 the immersion depth of the
vane, and distance from the bottom of the cup to the vane (Nguyen and Boger 1985).
With both vane and cup & bob geometries can be used with relatively small volumes
of slurry (as little as 200 ml).
Cone and plate geometries are used extensively with homogeneous fluids but are
unfortunately unsuitable for slurries, as the particles jam in the central region
between the cone and plate. The parallel-plate geometry, shown in Fig. 12.3d, has
the advantage that, like the cone and plate geometry, it requires only a small sample
size (as little as 10 ml). It is still necessary to ensure that the gap between the plates
meets the criterion set by Eq. (12.1). Unfortunately, unlike the cone and plate and
concentric cylinder arrangements, the shear stress is not constant throughout the
parallel-plate geometry, and data analysis is considerably more difficult (see, e.g.,
Macosko 1994).
With all the above geometries, rheological data is derived from the rotational
speed (N ) and the imposed or measured torque (T ). Note that these devices only give
meaningful results with viscometric flows (i.e., laminar flows) with known shear rate
profiles that will be discussed below. That is to say, the relationship between the
shear rate and the rotational speed must be known in order to yield useful results. At
sufficiently high shear rates (or stresses), secondary flows increase the torque for a
given shear rate, often in a quadratic manner, which may be misinterpreted as
turbulent flow. This instrument signature is not related to turbulence as experienced
in pipeline, pump, or mixing flows, and it cannot be used to give any insight into
these flows. Modern rotary rheometers are computer-controlled, allowing simple,
rapid analysis of controlled shear (or stress) behavior suitable for pipeline design, as
well as thixotropic analysis and elastic quantification.
The final geometry, Fig. 12.3e, is known as a capillary tube viscometer. This
comprises an accurately sized pipe through which the slurry is moved in a contin-
uous steady manner, either through pulseless pumps or from a pressurized vessel
1
A value of 10 is applicable for slurries for cv < 0.4cvmax for that material. However, visco-plastic or
yield stress slurries, common in the mineral industry, only have appreciable yield stresses at value of
cv approaching cvmax, where this equation is then applicable.
12.4 Rheology and Viscometry 375
discharging to atmosphere, or to another receiving vessel. The pipe (or tube) can
have either a vertical or a horizontal orientation. Small instrument capillary tube
viscometers with tube diameters of a few millimeters are used for homogeneous
fluids, while tubes used with slurries are typically 10 mm or more. Pipe test loops can
be successfully used as capillary tube viscometers; the only criterion being that they
must be operating in laminar flow to obtain any rheological data using the pressure
drops (Δp) and corresponding flow rates (Q).
Many engineers prefer to use a capillary tube viscometer for pipeline design,
claiming that because the flow is of the same form, spurious mismatches that might
occur when using rotational geometries are avoided. This logic is incorrect and many
reported differences are due to coarse particle interaction. For homogeneous fine
particle slurries, results obtained in rotary or capillary geometries are identical. For
coarser heterogeneous slurries, the coarser particles settle out of the measuring
region in rotary geometries, while in horizontal capillary geometries, they settle to
form sliding beds. This particle migration results in different and erroneous values in
both cases. Any differences observed between results obtained with well conducted
rotational geometry tests, and tests conducted in horizontal pipelines are due to the
migration of particles that are too coarse to form the non-Newtonian fluid (Pullum
et al. 2010). Unlike rotary viscometers, tube viscometers can also be operated in
turbulent flow, providing useful data for the design of turbulent flows in the full-
sized pipeline. Disadvantages of tube viscometers are that they require larger sample
sizes (typically more than 1000 l), temperature control is difficult, and they can only
be used for controlled shear rate (or shear stress) tests.
With all viscometers, two questions must be answered before any rheological
tests can be conducted:
What Are the Largest Particles That Should Be Used in the Viscometer?
Unfortunately, for most industrial suspensions, the upper size limit of particles that
form the internal structures of the carrier fluid, that is, those which may be consid-
ered to behave as a single phase when combined with the liquid, is surprisingly
small. Since the inter-particle forces are governed by a wide range of variables for
any particular system, a definitive upper particle diameter cannot be given; however,
evidence, obtained from experiments conducted to determine which particles may be
considered rheologically active and which are simply suspended by the carrier fluid,
suggests the upper limit lies somewhere between 0.5–20 μm (Pullum et al. 2015).
Consequently, most industrial slurries should be scalped before they are tested (i.e.,
the coarse fractions removed). In industrial applications it is recommended that
slurries should be scalped at around 40 μm, if possible, before viscometrical mea-
surements are made, while being cognizant of the effects such shearing may have on
the sample (i.e., how the slurry is passed through the sieve can impose a high shear
rate on the slurry which may alter the fluids rheology).
Over What Shear Rate Range Should the Tests Be Conducted?
Rheological tests must be conducted over a relevant range wherein the shear rate is
related to the representative velocities within the flow of interest divided by a
corresponding appropriate length scale (i.e., dU/dx). Extrapolation of results beyond
376 12 Testing and Instrumentation
ty
shear stress, t (Pa)
(Pa)
shear stress,
ty (a) (b)
Fig. 12.4 Rheograms for typical mineral slurries; (a) controlled strain or controlled stress
rheogram; (b) vane test rheogram
the tested range is not recommended. Some typical values of dU/dx for various
industrial processes are listed in Table 2.2. Also, because of possible thixotropic
behavior, the preparation of the sample before presentation to the rheometer should
mimic the process of interest, and often a pre-shearing of the sample at an appro-
priate shear rate, for a specified time, is required to attain a repeatable equilibrium
condition.
The two most common tests of direct use in hydraulic conveying and pumping are
the controlled strain (or controlled stress) test and the vane yield stress test. Typical
results for these two tests, displayed graphically as rheograms, are shown below.
In the first test, shown in Fig. 12.4a, either the shear rate (or shear stress,
depending upon the viscometer type) is varied across a range of values, and the
corresponding shear stresses (or shear rates) are measured. These values may be
selected manually or automatically varied across the range by computer control. In
both cases, it is necessary to ensure that sufficient time elapses to obtain steady-state
behavior before each data point is obtained. In this example, the relationship
between the shear stress and shear rate is clearly non-linear, and a minimum value
of applied shear stress before flow occurs, the yield stress (τy), is apparent. This yield
stress is sometimes known as the dynamic yield stress.
Figure 12.4b displays typical results from a vane test where a cruciform vane
(Fig. 12.3c) is inserted into a sample and very slowly rotated to obtain the trace
shown. Initially the slurry behaves like a solid, deforming elastically until it finally
yields at a maximum value, after which the fluid flows as the vane rotates. The
maximum stress value obtained is known as the static yield stress and is normally
considerably higher than the yield stress obtained in the previous test, under sheared
conditions. It is not uncommon for this static yield stress to be several multiples of
the dynamic yield stress.
12.5 Instrumentation 377
12.5 Instrumentation
Voltage across
electrodes
coils
measurement pipe
intensity of the magnetic field and the velocity of the material passing through the
meter. This signal is then processed to provide the bulk velocity (Q/A) through the
unit. The intensity of the magnetic field varies with distance from the signal
electrodes and across the pipe’s cross-section, so that units need to be calibrated
during manufacture, normally against a turbulent water velocity profile. In labora-
tory testing, it is common practice to calibrate a magnetic flow meter to an orifice
plate while running on water, before removing the orifice plate and commencing
slurry testing, or else using an end-of-line weigh tank for this purpose.
For non-Newtonian slurries undergoing laminar flow, the error introduced by the
change in flow regime is fortunately small, since shear-thinning slurries, typical of
the mineral and mining industries, have blunt laminar profiles. These profiles are
similar to a turbulent profile—and any turbulent profiles are also similar to that of
water (Heywood and Mehta 1999; Heywood et al. 1993). Considerable errors may
result, however, where there is a marked concentration gradient across the pipe, and
correspondingly a skewed velocity profile. For this reason, where stratified flows are
to be expected, the magnetic flow meter should be installed in a vertical pipe.
Heywood et al. (1993) also recommend that for slurry transport, measurements
should not be made below 40% of full scale, compared to the 10% limit usually
recommended for homogeneous fluids. When the solids being conveyed are mag-
netic (e.g., magnetite), the passage of the solids will disrupt the flow signal, typically
indicating a higher velocity than actual (Furlan et al. 2019).
Coriolis Flow Meter
Coriolis mass flow meters exploit the inertial forces that occur when a substance
moves through a rotating frame of reference. A schematic of a simple single-tube
meter is shown in Fig. 12.6, in which a tube is suspended and caused to vibrate.
Monitoring this motion are sensors that are mounted externally but close to the inlet
12.5 Instrumentation 379
and outlet of the tube. The tube’s resonant frequency will be a function of its mass,
including the substance inside it. Since the tube’s mass is constant, the frequency
then becomes a function of the mass, hence the density of the substance in the tube.
As the substance flows through the tube, forces due to the Coriolis acceleration twist
the tube such that the signals received by the inlet and outlet sensor are shifted
relative to each other. This phase change is proportional to the flow rate through the
tube. By monitoring the frequency and phase shift of the signals from the two
sensors, the density of the substance in the pipe, the volumetric flow rate, and
consequently the mass flow rate through the pipe can be monitored to a high degree
of accuracy. For this reason, Coriolis meters are often found in the oil and gas
industry, where they are used for fiscal transfer measurements.
Industrial instruments take several forms, using either single, twin, or multiple
parallel tubes in pipe diameters ranging from 6 to 500 mm. For the instrument to be
sensitive to both density and flow rate, the tube (or tubes) must be as light as
possible, and hence thin. Consequently, these meters are not suitable for slurries
that are strongly erosive. Even for mildly erosive slurries, special materials may be
needed. Where multiple tubes are used, their diameter is reduced compared to the
conveying pipeline. This may increase velocity and wear, and will reduce the
maximum practical coarse particle size. Although units are available for relatively
large pipe sizes, the instrument and installation costs can become very high as the
pipe size increases. Furthermore, for suspension flows, the tube’s vibration induces
virtual mass effects that can produce errors of up to 15% of the solids mass flow at
volume concentrations of 40% (Basse 2014). Despite all of this, the ability to
simultaneously measure the density and the flow rate, especially where air entrain-
ment is present, provides an attractive alternative to other instrumentation for
moderate flows of fine particle slurries.
Bend Meter
A simple device that lends itself readily to slurry use, the bend meter is shown
schematically in Fig. 12.7. As with an orifice plate or venturi, flow rate is correlated
to a pressure difference caused by a change in the momentum of a flowing fluid. In
the bend meter, flow rate is proportional to the square root of the pressure difference
measured across the bend (corrected for any difference in elevation between the two
380 12 Testing and Instrumentation
45⁰
Qm
measurement points). To a first approximation (Bean 1971) the slurry flow rate for a
90° bend, with pressures measured at the 45° positions as shown in Fig. 12.7, is:
rffiffiffiffiffiffiffiffiffi
πD2 RΔp
Q≈ ð12:2Þ
4 Dρm
where D is the pipe diameter, R is the bend radius, and Δp is the pressure difference
from the outside to the inside of the bend. While Eq. (12.2) may be used for bend
meter pipe and pressure transducer sizing, it is advisable to calibrate any specific
bend meter against a more repeatable instrument, such as an orifice plate or weigh
tank, due to variability in construction. Because of the small pressure differential,
“wet” tappings (Sect. 12.5.2) are normally required for acceptable accuracy; thus,
bend meters are usually restricted to the laboratory environment, where they are
sometimes used to provide a backup check of the primary flow instrument during
slurry testing.
A particular advantage of this technique is that it can be applied to an existing
bend and requires no additional disturbance to the slurry. The precautions explained
above must be used to ensure that the pressure readings are reliable. This type of
device becomes less reliable if the flow is stratified. It is therefore preferable to apply
the measurement at a bend where the entering flow is vertical (i.e., at the top of a
“riser” or the bottom of a “downcomer”). Experience at the GIW Hydraulic Labo-
ratory has shown that, provided these precautions are taken, the bend meter is
reliable and gives reproducible measurements (Addie 1981).
12.5 Instrumentation 381
Acoustic Methods
Since the development of neural net analysis, there have been several attempts to
predict volumetric and/or mass flow rate through systems by listening to the sounds
generated by inter-particle and particle-wall collisions. However, as these networks
need to be trained by sample data, the instruments are usually restricted for use only
in the original equipment where they were trained. The gathering of very large
amounts of training data required for modern deep learning techniques for use in
developing suitable artificial intelligence (AI) instrumentation seems, at the time of
writing, impractical (Fig. 12.8).
Rather than “listening” to the stochastic sounds generated in the system, flow
meters have been developed (e.g., Gysling et al. 2005) to exploit the fact that
turbulent structures in the flow are convected along the pipe by the mean velocity
of the flow. The local pressure variations, produced by these eddies, can be sensed by
strain or acoustic sensors applied to the outer wall of the pipe. A linear array of these
sensors, fitted along the pipe, monitor the passing eddy structures to obtain the mean
velocity, and hence the flow rate through the pipe. For pseudo-homogeneous flow,
where the slip velocity between the phases is small, this will also correspond to the
mixture flow rate of the slurry. These meters can provide accuracy similar to that of a
magnetic flow meter under laboratory conditions, but they have the advantage that
sensing hardware is not exposed to the slurry and not affected by any magnetism
within the slurry.
By “listening” to the acoustically generated signals within the pipe in a similar
manner, albeit at a much different propagation velocity, the acoustic velocity in the
slurry can also be determined. This characteristic velocity is a strong function of air
concentration and thus can be used to estimate the volume of air entrainment within
the slurry.
382 12 Testing and Instrumentation
12.5.2 Pressure
to pressure pot
to pressure pot or transducer
or transducer
full bore transparent
ball valve tubing
min.
3mm
purge water hole
~60
⁰
bleed from to
to tapping transducer
transducer
from
tapping
drain
drain
Fig. 12.9 Pressure tappings and pressure pots suitable for slurry flows
Unfortunately, secondary flows within the blind leg of the resulting large “tee” will
alter the pressure transmitted through the diaphragm. For simple static pressure
measurements, such errors may be small compared with the pressure inside the
pipe, but where two static readings are used to calculate a pressure gradient, the
errors may be significant. Furthermore, the magnitude of the error will change with
flow rate.
Establishing a pressure gradient using different static transducers usually requires
differencing two similarly large pressures and is always less accurate than using a
differential pressure transducer. Care must still be exercised, however, when using
diaphragm isolators in conjunction with differential transducers. The tapping lines
384 12 Testing and Instrumentation
between the differential pressure transducers and isolating diaphragms are a closed
system and act as a constant volume thermometer, so temperature differences
between the two tapping lines on each side of a transducer (perhaps due to differ-
ences in ambient or solar radiation) may bias the pressure reading.
Density Flasks
The density of samples taken from the system can be readily measured by weighing
flasks of known volumes filled with the slurry. The same instruments can be used to
measure the pure liquid density.
Figure 12.10 shows two different density flasks suitable for use on-site. The first
(a) is the familiar Marcy gauge, which is suitable for fine particle slurries. After the
weight of the flask is tared out, the flask is filled with a well-mixed sample of the
slurry to just above the vents that are located in the side of the flask and define the
flask volume. After the excess slurry is carefully washed from the outside of the
flask, the density and concentration of the slurry may be read directly from a multi-
scaled spring balance using the scale corresponding to the solids density. For rapidly
settling suspensions, this method will slightly overestimate the density, since a
disproportionate amount of the supernatant fluid will flow out of the drainage vents.
For coarse-particle suspensions, especially those with high concentrations, a
different technique can be employed. In Fig. 12.10b, a transparent jar with a lid is
tared out on a suitable electronic scale. The lid is then removed, and the jar is almost
completely filled with a well-mixed suspension sample as shown. The lid is then
central meniscus
removable lid gauge
filling vent
extra water
layer
transparent jar
coarse slurry
sample
drain vents
Fig. 12.10 Density flasks suitable for slurries (a) a Marcy gauge and (b) a flask suitable for coarse
suspensions
12.5 Instrumentation 385
replaced, and the weight of the sample (Ws) recorded. Using a pipette, a layer of
water is then carefully introduced into the flask through the filling vent until the
meniscus of the water attaches to the end of the sharp gauge pin. The weight of the
sample and water (Wt) is also recorded. Assuming the volume of the flask (as defined
by the jar’s internal diameter and meniscus gauge height) is V, then the density of the
sample is given by:
Ws
ρm = ð12:3Þ
Wt - Ws
V-
ρw
where ρw is the density of water. Provided the meniscus gauge is placed centrally in
the lid, there is no need to accurately level the flask for satisfactory operation.
Nuclear Densitometer (Gamma Gauge)
Slurry density can be measured using a meter that relies on the attenuation of some
form of radiation passed through the slurry. Gamma rays are most commonly used,
and this type of instrument is sometimes referred to as a nuclear densitometer
(Fig. 12.11). The radiation falling on the detector is proportional to the source’s
strength and the mass of the matter between the source and the detector; conse-
quently, the parameter measured is the mean density along the radiation path, which
indicates the in situ density rather than the delivered value. This must be remembered
if using the output from this instrument to calculate solids throughput. Calibration
will be affected by the thickness and materials of the intervening pipe wall (which
may change over time), as well as the condition of the source (which will decay with
time).
In stratified flow, the measurement will be affected by the location of the gamma
ray beam within the slurry. As with other instruments covered in this chapter,
accuracy with settling slurries is therefore optimized when the unit is installed on a
straight, vertical section of pipe. In horizontal installations, the beam is often passed
through the pipe at a 45° angle in an attempt to select an averaged concentration
distribution, avoiding the lowest and highest concentration areas at the top and
bottom of the pipe. This may provide sufficient accuracy for process control, but
will still present significant errors which vary with the flow velocity and degree of
stratification.
Due to occupational safety and health concerns associated with using radioactive
materials in the workplace, installation of such devices has become increasingly
difficult, and alternative methods to monitor density have been sought.
Direct Weighing Methods
A simple method to measure the density of the slurry directly is to weigh a horizontal
section of pipe of known volume—or to otherwise determine its weight via a
secondary measurement, such as pipe deflection. As with the nuclear densitometer,
these instruments measure the in-situ (rather than the delivered) density. Further-
more, since they must be installed horizontally, the effects of stratification and
deposition that increase the difference between the in-situ and delivered density
must be considered when pumping heterogeneous suspensions (Fig. 12.12).
There are also mechanical challenges, such as the need to provide flexible joints
between the weighing section and the normal piping—joints that will not entrap
solids and become rigid will suitably decouple axial thrust from the load measure-
ment, and will cope with vibrational loads inherent in any installation. The weight of
the pipe also decreases the sensitivity of the slurry mass measurement. At the time of
this writing, several devices have been shown to overcome these obstacles by
encasing a flexible pipe suspended inside a pressure jacket and measuring the
mass of the comparatively light flexible tube via a load cell, the output of which is
suitably filtered to remove extraneous vibrational signals. However, user experi-
ences with such devices have been mixed.
By averaging the differential pressure drops over vertical up and down sections
within the loop, the friction losses cancel and the in-situ concentration of the slurry
may be calculated according to Eq. (7.10), which is repeated below. For best
accuracy (based on experience in the GIW Hydraulic Laboratory), the pressure
taps should begin at least ten pipe diameters downstream, and one diameter
upstream, from the adjacent elbows. The distance between taps should also be at
least ten pipe diameters.
Δp12,man þ Δp43,man
C vi = ð7:10Þ
2ðSs - Sf Þρw gΔz
388 12 Testing and Instrumentation
While this instrument measures the average in-situ concentration between the two
vertical sections, it has been shown that this value approximates the delivered
concentration (Cvd), even with larger solids having significant settling velocities,
and where measurable differences in density are seen between the instrument’s up-
and down-legs. This result is obtained because the effects of solids slip in the two
vertical sections largely cancel each other in the average (with regard to concentra-
tion), and also because the associated friction losses are less sensitive to concentra-
tion, since particle-to-wall collisions are minimized in vertical flow. The U-loop also
often provides a convenient location for the vertical mounting of a magnetic or
acoustic flow meter.
Unfortunately, because the instrument relies on the resolution of a small differ-
ence between two independent pressure measurements, “wet” tappings with associ-
ated settling pots and frequent cleaning are required; thus, U-loop meters are
normally restricted to the laboratory environment. Attempts by the GIW Hydraulic
Laboratory to apply isolated pressure transducers suitable for a production installa-
tion have not been successful. The logistics of installing a U-loop in the field are also
not trivial, and this is rarely attempted for other than temporary research purposes
(Fig. 12.13).
Fig. 12.13 Inverted U-loop used in a field application. (GIW Hydraulic Lab)
12.5 Instrumentation 389
Deposition limit velocity can be, and often is, visually observed through a transpar-
ent section during laboratory testing. Since the manner of deposition can vary
dramatically with particle size, concentration, and carrier fluid properties, visual
observation provides additional information to the researcher beyond the simple fact
of whether or not a stationary deposit exists. However, this is often not possible in
field installations where a transparent pipe section is impractical, or where the slurry
is rendered opaque by a high concentration of fine particles. For these situations,
instruments have been devised that detect the presence of a stationary bed via
thermal measurements (Ercolani et al. 1979; Ilgner 2017). The early sensors simply
monitored electrical or thermal activity close to the probe as an indication of particle
movement, to provide a rapid indication of a stationary deposit. Later devices
delivered a fixed thermal load into the pipe wall while measuring the resulting
temperature (e.g., Ilgner 2014). If a stationary bed is present on the inside of the
pipe adjacent to a sensor, the thermal convection is significantly reduced, resulting in
an increase in the pipe wall temperature. By monitoring the temperature difference
between the top and the bottom of the pipe, an observer can then determine whether
or not a stationary bed is present. If more sensors are added along the sidewalls of the
pipe, the observer can also infer the height of the stationary bed. These instruments
must be calibrated to account for various effects, such as sunlight and shade, slurry
type, and pipe wall properties; however, they have the advantage of being
non-invasive and relatively simple to install and operate.
The measurement of pump input power provides a useful diagnostic tool in trouble-
shooting and operation, for reasons already considered in Chaps. 8 and 11. For direct
measurement of pump input power, it is preferable to use a separate transducer (Sect.
12.3), but these require extra couplings and a space in the drive train to accommodate
them; therefore, they are usually used only in laboratory tests or in specially arranged
field tests.
For pumps with electric drives, pump input power can be estimated from the
delivered voltage (E) and current (I). If a three-phase induction motor is used, the
motor power can be estimated as:
pffiffiffi
P = 3EIη cos θ ð12:4Þ
where θ is the phase angle between voltage and current, and η is the drive efficiency,
which is equal to the product of the individual efficiencies of all drive components
between the electrical supply and pump shaft. For a fully loaded motor, cosθ can be
estimated as 0.8. Two-element type wattmeters, using current and potential
390 12 Testing and Instrumentation
Various methods are available to measure the internal velocity profiles of one or
more of the phases within a flow stream. Generally, the erosive nature of slurry flows
prevents the use of probes (e.g., Pitot tubes, hot-film probes, and ballistic probes),
and their presence can also distort the flow being measured. Consequently, only
non-intrusive techniques will be discussed here.
Ultrasonic Doppler Velocity Profilers (UDVP)
In this technique, an acoustically coupled ultrasonic transducer sends a series of
ultrasonic bursts along an inclined path as shown in the top pane of Fig. 12.14a.
Scatter centers (very small particles) in the slurry reflect these emissions, the
frequency of which will be shifted from the original frequency through the Doppler
Fig. 12.14 UDVP. (a) schematic showing installation, signal, and derived velocity profile (Takeda
1995) and (b) experimental velocity profiles
12.5 Instrumentation 391
laser
measuring region
showing fringe pattern
receiving lens &
photo detector for
particle phase
(a) (b)
Fig. 12.15 LDV technique (a) Components required for a single-component back scatter system,
(b) Velocity profiles obtained in a 25% matched refractive index slurry. (Wildman et al. 1992)
shift. Knowing the time of detection of these reflections and the acoustic velocity in
the slurry enables the location along the path to be determined, and the Doppler shift
and geometry enable the axial velocity of the scatter center(s) that reflected the signal
to be calculated. A velocity profile, shown in the lower pane of Fig. 12.14a, can thus
be constructed. Examples of velocity profiles obtained in a sand slurry are shown in
Fig. 12.14b. The penetration of the acoustic beam is a function of the beam’s
frequency, the particle density, and concentration of the slurry. If the mean spatial
concentration of solids is high, then the ultrasonic techniques can be used to scan the
entire flow depth only if low-density particles are used for the flow testing (e.g.,
Matoušek et al. 2015).
Laser Doppler Velocimetry (LDV)
The gold standard for fluid flow measurement, LDV provides accurate measure-
ments of laminar or turbulent flow velocities in one to three dimensions. The usual
technique requires shining pairs of laser beams into the flow that is seeded with small
scatter centers or tracer particles. The beams intersect in the flow, defining a small
measuring volume where they cross. This measuring volume is used to interrogate
the flow field. By using beams of different colors (e.g., the various lines of an argon
ion laser) and arranging these beams in different planes, the various components of
the velocity in the measuring volume can be obtained by color separation. Velocities
in cyclical or highly turbulent flows can also be resolved using frequency-shifting
techniques. A common geometry for a single-component system is given in
Fig. 12.15a.
Since the laser needs to penetrate the flow, it must be optically transparent. For
flows of very dilute concentration, both phases can be measured by using collecting
optics placed in full backscatter to measure the fluid phases, as shown, and at 32° to
measure the solids phase. For higher concentration slurries, matched refractive index
materials are needed (see section below); but with care, highly concentrated suspen-
sions can be probed using this method as illustrated in Fig. 12.15b.
392 12 Testing and Instrumentation
1
0.8
0.6
0.4
0.2
u/Umax
r/R
0
-0.2 0.0 0.5 1.0 1.5
-0.4
-0.6
solids
-0.8 fluid
-1
(a) (b)
Fig. 12.16 PIV, (a) components for single phase system, (b) simultaneous particle and fluid
velocity profiles
In the previous edition of this book, we speculated on the potential for applying
computer-aided tomography techniques (CAT scanners), then in use for medical
purposes, to the future of pipeline research and monitoring. Such is the advance of
technology that two- and three-dimensional scanning of these systems is now
commonplace in research laboratories and can even be found in some industrial
applications (see e.g., Wang 2022).
Gamma Ray Profiler
A gamma ray-based method and technique is relatively frequently used to measure
the spatial concentration of solids in solid-liquid flows through pipes. Typically, a
12.5 Instrumentation 393
Fig. 12.17 Gamma ray measuring unit for laboratory pipe. (1) gamma ray source, (2a) collimator at
source, (2b) collimator at detector, (3) scintillator, (4) photomultiplier, (5) digital analyzer, (6) lead
shielding of detector, (7) pipe cross-section, (8) linear positioning drive, (9) radial positioning drive.
(Krupička and Matoušek 2014)
measuring system is composed of one source of a gamma ray beam on one side of a
pipe cross-section, and one detector on the opposite side. In field pipelines, the
measuring unit is typically mounted to a vertical section of the pipeline and a
measuring gamma ray beam is directed to the center of a pipe cross-section. This
way, the measured concentration can be interpreted as the mean spatial concentration
in the cross-section. In laboratory pipe loops, the unit is mounted to a measuring
section of the pipe (either horizontal or inclined) and operated so that it produces a
profile of chord-averaged concentrations in the pipe cross-section (e.g., Gillies 1993;
Krupička 2014; Matoušek 1997; Pugh 1995). The position of the measuring beam
can be varied across the pipe cross-section (positioning drive No. 8 in Fig. 12.17),
and the beam is often collimated to increase the spatial resolution of the measured
local concentration.
Electrical Resistance Tomography (ERT)
ERT is by far the most common form of tomography used in laboratories, as it is
relatively inexpensive, safe, and rapid. The technique involves passing a high-
frequency current through pairs of internal surface-mounted electrodes, and measur-
ing the voltages on several other electrodes similarly mounted in a transverse plane
or electrode ring (Fig. 12.18a). The electrical response is a function of the conduc-
tivity in the pipe’s cross-section and is used as the boundary conditions in a
computational reconstruction of the conductivity map within the pipe in the plane
of the electrodes. The local conductivity is related to the local solids concentration
through the Maxwell equation (Maxwell 1873), and so acts as a proxy for the
concentration map.
394 12 Testing and Instrumentation
s1
electrode pairs
excited and
s2
I
high frequency
(a) (b)
Fig. 12.18 ERT system (a) operational schematic (b) typical laboratory and industrial electrode
rings. (Lower image courtesy: Industrial Tomography Systems Ltd.)
Carrier
Fig. 12.19 Typical ERT results showing particle settling in a 400 Pa tailings. (Graham et al. 2017)
Fig. 12.20 Magnetic resonance imaging of laminar flow in a hybrid slurry (a) concentration map,
(b) axial velocity map, (c) concentration and velocity profiles on the vertical central plane. (Pullum
and Graham 2000)
measurements to be made using just the carrier fluid alone, an inconvenience in the
laboratory and normally an impossibility on-site.
By taking an area average of the solids concentration over the measurement
plane, the in-situ solids concentration can be measured with ERT. As such, com-
mercial ERT units are an attractive alternative to nuclear densometers for field slurry
density measurements, as they do not require training and maintenance, nor do they
pose any of the occupational health and safety issues associated with devices
containing a radioactive source.
Very fast systems are now available, some of which even capture gross charac-
teristics of turbulent puff data. By combining the output from two or more electrode
rings (planes) the slurry’s velocity map can also be measured (e.g., Lucas et al. 1999;
Faraj and Wang 2012).
Magnetic Resonance Imaging (MRI)
MRI has the advantage of being sensitive to both flow and local density variation. As
a result, it can produce both high-resolution three-dimensional concentration and
velocity maps, as shown in Fig. 12.20.
However, this technique is expensive and requires custom installation, and any
magnetic material in the experimental equipment (including the material being
pumped) will produce image artifacts. One of the authors has experience of a fly
ash slurry that contained enough residual iron (from the pulverizers used to produce
the coal fuel), to exclude it from testing. Furthermore, like a super magnetic
separator, the MRI would have retained all of the residual iron.
396 12 Testing and Instrumentation
array of 17 detectors
and pre-amplifiers
lead collimator
plates
measuring volume
80 dia x 10mm deep
Fig. 12.21 Gamma ray scanner (a) schematic (Hjertaker 1998), (b) unit installed in the SRC’s Pipe
flow technology center, Saskatoon. (Courtesy Saskatchewan Research Council)
References
G.R. Addie, Slurry pump and pipeline performance testing at Georgia Iron Works hydraulic
laboratory. Presented at the 6th international technical conference on slurry transportation,
Las Vegas, 1981
ANSI/HI 14.6, American National Standard for rotodynamic pumps for hydraulic performance
acceptance tests, Appendix K. Hydraulic Institute, Parsippany, 2022
H.A. Barnes, Measuring the viscosity of large-particle (and flocculated) suspensions – A note on the
necessary gap size of rotational viscometers. J. Non-Newtonian Fluid Mech. 94, 213–217
(2000)
H.A. Barnes, J.F. Hutton, K. Walters, An Introduction to Rheology (Elsevier Science BV, Amster-
dam, 1989) ISBN 0-444-87140-3
References 397
N.T. Basse, A review of the theory of coriolis flowmeter measurement errors due to entrained
particles. Flow Meas. Instrum. 37, 107–118 (2014)
H.S. Bean, Fluid Meters Their Theory and Application (ASME, New York, 1971)
D. Ercolani, F. Ferrini, V. Arrigoni, Electric and thermic probes for measuring the limit deposit
velocity. Presented at the 6th international conference on the hydraulic transport of solids in
pipes, Canterbury, 1979
Y. Faraj, M. Wang, ERT investigation on horizontal and vertical counter-gravity slurry flow in
pipelines. Procedia. Eng. 42, 588–606 (2012)
J. Furlan, R. Visintainer, A. Sellgren, V. Matoušek, L. Pullum, Pipe loop testing of a mixture
containing fine dense solids with magnetic properties. Presented at the 19th international
conference on transport and sedimentation of solid particles, Cape Town, 2019
R.G. Gillies, Pipeline flow of coarse particle slurries. PhD Thesis, University of Saskatchewan,
Saskatoon, 1993
L. Graham, L. Pullum, A. Chyrss, Absolute 2.5d electrical resistance tomography in pipe systems.
Presented at the 20th international conference on hydrotransport, Melbourne, 2017
D.L. Gysling, D.H. Loose, A.M. van der Spek, Clamp-on, sonar-based volumetric flow rate and gas
volume fraction measurement for industrial applications. Presented at the 13th international flow
measurment conference FLOMEKO, Peebles, 2005
S.A. Hashemi, R.B. Spelay, R.S. Sanders, B.T. Hjertaker, A novel method to improve electrical
resistance tomography measurements on slurries containing clay. Flow Meas. Instrum. 80, 1–11
(2021)
N.I. Heywood, K.B. Mehta, Performance evaluation of eight electromagnetic flowmeters with fine
sand slurries. Presented at the 14th international conference on slurry handling and pipeline
transport, Maastricht, 1999
N.I. Heywood, K.B. Mehta, D. Poplar, Evaluation of seven commercially available electromagnetic
flowmeters in solid/liquid mixtures. Presented at the 12th international conference on slurry
handling and pipeline transport, Brugge, 1993
B.T. Hjertaker, Static characterization of a dual sensor flow imaging system. Meas. Sci. Technol. 9,
183–191 (1998)
B.T. Hjertaker, R. Maad, E. Schuster, O.A. Almås, G.A. Johansen, A Data Acquisition and Control
System for High-Speed Gamma-Ray Tomography. Meas. Sci. Tech. 19, 094012, 1–9 (2008)
H.J. Ilgner, Novel instrumentation to detect sliding and erractic bed load motion. Presented at the
19th international conference on Hydrotransport, Golden, 2014
H. Ilgner, Non-invasive detection of sedimentation and its removal in industrial pipelines. Presented
at the 18th international conference on transport and sedimentation of solid particles,
Prague, 2017
F. Irgens, Rheology and Non-Newtonian Fluids (Springer, New York, 2014)
ISO 9906, Rotodynamic Pumps – Hydraulic Performance Acceptance Tests – Grades 1, 2 and 3
(International Organization for Standardization, Geneva, 2012)
J. Krupička, V. Matoušek, Gamma-ray-based measurement of concentration distribution in pipe
flow of settling slurry: Vertical profiles and tomographic maps. J. Hydrol. Hydromech. 62,
126–132 (2014)
G. Lucas, J. Cory, R.C. Waterfall, W.W. Loh, F.J. Dickin, Measurement of the solids volume
fraction and velocity distributions in solid-liquid flows uisng dual-plane electrical resistance
tomography. Flow Meas. Instrum. 10, 249–258 (1999)
C.W. Macosko, Rheology: Principles, Measurements and Applications (VCH Publications,
New York, 1994)
V. Matoušek, Flow mechanism of sand-water mixtures in pipelines. Dissertation, Delft University
of Technology, Delft, The Netherlands, 1997
V. Matoušek, J. Krupička, T. Picek, Š. Zrostlík, Experimental investigation of internal structure of
open-channel flow with intense transport of sediment. J. Hydrol. Hydromech. 63, 318–326
(2015)
J.C. Maxwell, A Treatise on Electricity and Magnetism, vol 1 (University of Oxford, 1873)
398 12 Testing and Instrumentation
Q.D. Nguyen, D.V. Boger, Direct yield stress measurement with the vane method. J. Rheol. 29,
335–347 (1985)
F.J. Pugh, Bed-load velocity and concentration profiles in high shear stress flows. PhD Thesis,
Queen’s University, Canada, 1995
L. Pullum, L.J.W. Graham, The use of magnetic resonance imaging (MRI) to probe complex hybrid
suspension flows. Presented at the 10th international conference on transport and sedimentation
of solid particles, Wroclaw, 2000
L. Pullum, P. Slatter, L. Graham, A. Chryss, Are tube viscometer data valid for suspension flows?
Korea Aust Rheol. J. 22, 225–230 (2010)
L. Pullum, A. Chryss, L. Graham, V. Matoušek, V. Pĕník, Modelling turbulent transport of solids in
non-newtonian carrier fluids applicable to tailings disposal. Presented at the 17th international
conference on transport and sedimentation of solid particles, Delft, 2015
Y. Takeda, Velocity profile measurement by ultrasonic Doppler method. Exp. Therm. Fluid. Sci.
10, 444–453 (1995)
M. Wang (ed.), Industrial Tomography Systems and Applications, 2nd edn. (Elsevier, Amsterdam,
2022)
D.J. Wildman, J.M. Ekmann, J.R. Kadambi, R.C. Chen, Study of the flow properties of slurries
using the refractive index matching technique LDV. Powder Tech. 73, 211–218 (1992)
Chapter 13
Erosive Wear
13.1 Introduction
Wear is a fact of everyday life. Almost everything we use eventually “wears out” due
to the various types of physical contact to which the surfaces of these objects are
exposed. The mechanisms of wear can take many forms. “Two bodies” may be
involved, as when a wheel rubs on a clean hard floor; or there may be “three bodies,”
as when sand is present between the wheel and floor. Bodies may be constrained, as
with the floor, held with force against another object, as with the wheel, or freely
moving between objects, as with the sand.
Much of the wear experienced by slurry handling equipment is two-body wear
caused by freely moving particles within a fluid flow, sliding or impinging against a
solid surface. This type of wear is often called erosive wear and will be this chapter’s
primary focus.
The useful life of most slurry transport equipment is limited by the erosive wear
of the wetted passages. Obvious factors in wear performance are the properties of the
slurry solids and the materials from which the wetted passages are constructed.
However, the wetted passages are also the “working” parts of the equipment,
meaning that their purpose is to control the flow of the slurry as part of the industrial
process. This may be for guiding the flow (pipelines and launders), adding energy
(centrifugal pumps), or modifying the slurry properties (hydrocyclones, flotation
cells, thickeners, etc.). There may be a strong dependence between the hydraulic
performance or operating conditions of the equipment and its resulting wear perfor-
mance. Conversely, wear will often alter the hydraulic performance of the equipment
itself, over time. Figure 13.1 shows, side by side, a new and worn impeller of the
same design. Note that the worn impeller was still achieving the required process
conditions for the pump at the end of its life, albeit with some alteration in
performance (increased speed) and some reduction in efficiency (increased power).
In the preceding chapters, we have seen that the mechanisms of slurry flow are
complex and not always easily understood or predicted. The small-scale interactions
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 399
R. Visintainer et al., Slurry Transport Using Centrifugal Pumps,
https://doi.org/10.1007/978-3-031-25440-6_13
400 13 Erosive Wear
between billions of slurry particles and the wetted surfaces of slurry equipment add
another layer of complexity when it comes to wear analysis and prediction. As a
result, a good deal of wear performance evaluation has occurred post facto, when the
system is already in operation and it is possible to correlate wear performance to
equipment design and average operating conditions. A body of experience and
insight gathered by this method has accumulated over time, and many of the design
strategies for improving wear performance in slurry systems are based on this
experience.
Recent decades have seen the introduction of more rigorous approaches to wear
performance evaluation, allowing for progress to be made in a more systematic way.
These include standardized laboratory tests for ranking slurry abrasivity and material
wear resistance, and electron microscopy for providing close examination of the
micro-mechanisms of wear. The numerical modeling of slurry flow has also seen
considerable development as more powerful computers have become widely avail-
able and numerical techniques more refined. This chapter introduces the mechanisms
of erosive wear, the various materials used in high wear applications, and some of
the more useful methods for erosive wear testing. This leads to an overview of wear
modeling techniques and a discussion of the operational and design considerations
that affect wear performance in slurry pumps and pipelines. A particular focus is
given to the centrifugal slurry pump. Since the pump’s primary function is to add
energy to the slurry, the fluid velocities and accelerations within its wetted passages
are high. This results in high wear rates, making the slurry pump one of the most
maintenance-intensive pieces of equipment in the slurry processing operation. To
13.2 Mechanisms of Erosive Wear 401
this end, a detailed overview is provided of the insight that experience and numerical
modeling have provided into the wear performance of slurry pump wet-end
components.
The mechanisms of erosive wear span a continuum from single-particle impact to the
sliding of a compact bed of slurry particles, here referred to as sliding-abrasion wear.
The majority of actual slurry erosion processes fall somewhere between, where a
dense mixture of solids and fluid interact with the wetted surfaces (and each other) in
a combination of low-angle impact and sliding-abrasion wear. However, for the
purposes of testing and modeling, it remains useful to examine the extremes of pure
impact wear and sliding-abrasion.
The sliding-abrasion mode of wear typically involves a bed of contact-load
particles bearing against a surface and moving tangent to it, as illustrated in
Fig. 13.2. In pipelines, the stress normal to the surface is usually caused by gravity,
as shown in the analysis of contact load presented earlier in this book. As described
in Chaps. 3 and 4, the submerged weight of particles that are not suspended by the
fluid must be carried by intergranular contacts. The analysis of the motion of these
contact-load (bed-load) solids in pipes has been developed over many years (see
discussion of the two-layer model in Sect. 4.3.1). For sliding-abrasion, the erosion
rate depends on the properties of the particles and wear surfaces, the normal stress,
and the relative velocity.
The normal stress is increased when the flow streamlines are curved, as in a pipe
elbow. In this case, there is a centrifugal acceleration, equal to u2/r, where u is the
local velocity and r is the radius of curvature of the local streamlines. This acceler-
ation can often be much greater than that of gravity, producing a commensurate
increase in the normal stress between the moving contact-load solids and the wall
material, and hence a greatly increased rate of sliding-abrasion wear. This type of
behavior can cause elbows to wear through when adjacent straight pipe sections of
the same material have hardly been affected by erosion. It is also very important in
centrifugal pump casings where sliding-abrasion tends to dominate in most areas.
The second type of wear is the particle impact mode, which occurs where
individual particles strike the wearing surface at an angle, despite the fact that the
fluid component of the slurry necessarily conforms to the surface (see Fig. 13.3).
Removal of material over time occurs through small-scale deformation, cutting,
fatigue cracking, or a combination of these, and it thus depends on the properties
of both the wearing surface and the particles. Ductile materials tend to exhibit
erosion primarily by deformation and cutting, with the specific type depending on
the angularity of the eroding particles. Brittle or hardened materials tend to exhibit
fatigue cracking erosion under repeated particle impacts. For a given slurry, the
erosion rate depends on the properties of the wearing surface: hardness, ductility,
toughness, and microstructure. The mean impact velocity and mean angle of impact
of the solids are also important variables, as are particle characteristics such as size,
shape and hardness, and the concentration of solids near the surface.
Particle impact erosion occurs because the trajectories of the individual particles
do not follow the streamlines of the flow. This behavior is important where the
13.2 Mechanisms of Erosive Wear 403
solidification and a low heat conductivity, requiring careful mold design and control
of stresses during heat treatment. Obtaining a material strength suitable for construc-
tion of slurry equipment components, while also having an optimized carbide
distribution for wear resistance, requires specialized alloy additions and processing
parameters. This requirement applies especially to larger, complex shapes, such as
the typical centrifugal slurry pump casing or impeller.
Strategies for addressing these challenges vary widely, and well-defined stan-
dards for commercial white iron metallurgy do not exist. Rather, individual
foundries often develop proprietary grades to optimize the performance of their
products. Challenges notwithstanding, white irons provide an excellent compromise
between wear resistance, toughness, manufacturability, and cost, and they are widely
employed in all manner of slurry system machinery, including slurry pumps;
grinding, crushing, and conveying equipment; pipeline fittings; and valves.
Elastomers
For slurries without coarse particles, elastomers such as rubber and polyurethane can
provide wear resistance exceeding that of hardened metals. These are long-chain
polymers with a cross-linked molecular structure that allows them to sustain large
elastic deformations without damage. They resist wear by their ability to deform and
rebound to their original shape without absorbing much energy; the energy is
returned instead to the erosive particles. This property, often referred to as the
resilience of the elastomer, together with the particle properties and type of erosion
(impact or sliding), determine the performance of the elastomer. Particle size is an
important factor, and the average (d50) particle size must generally be limited for
elastomers to be successful. A particle top-size limitation also applies, including
tramp (extraneous material such as occasional rocks, bolts, or pieces broken from the
upstream process equipment), which can quickly damage elastomer-lined parts and
lead to sudden failure. The crossover point in particle size where elastomers become
406 13 Erosive Wear
16
14
12
intermediate metal preferred
10
dTOPSIZE (mm)
4
rubber preferred
2
0
0 1 2 3
d50 (mm)
Fig. 13.6 Approximate boundary for successful application of rubber-lined, centrifugal slurry
pumps
more effective than hardened metals is not a clearly defined boundary, but rather a
range that will vary with the properties of the elastomer, the slurry, the fluid flow
properties, and the type of equipment. A general guide for centrifugal slurry pump
applications can be seen in Fig. 13.6. Other types of equipment, where the velocities
and accelerations are lower than inside a slurry pump, may allow larger solids.
Elastomers are subject to other limits that must be considered. They typically
have lower service temperature limits than metals, and are subject to different
sensitivities to chemicals within the slurry flow. They are also subject to limitations
in the velocity and energy of the slurry flow, which can cause internal vibrations in
the elastomer. These can generate heat at a higher rate than it can be dissipated,
leading to a breakdown of the material often referred to as hysteresis damage, after
the non-linear behavior of the elastic properties that lead to heat generation (see
Fig. 13.7). In centrifugal slurry pumps, this damage can affect both impellers and
liners, and it imposes a limit on the head per impeller stage to about 40 m for a
typical rubber lining material.
There are many formulations of elastomers and more are developed every year.
However, for slurry duty, natural rubbers, neoprene rubbers, and polyurethanes
account for most of the current use.
• Natural rubber: Compositions of natural rubber, manufactured from latex poly-
mer obtained from the sap of the rubber tree, remain the most common elastomers
in use for slurry service due to their combination of outstanding resilience, good
resistance to many chemicals, and low cost. They are produced by vulcanization,
a process during which the raw polymer is mixed with various chemicals, and a
cross-linked molecular structure is created under the application of heat and
pressure. Natural rubbers have a relatively low-temperature limit for wear service
13.3 Wear-Resistant Materials 407
(about 65–80 °C, depending upon their formulation) and poor resistance to even
small traces of petroleum-based chemicals; however, when these limitations are
not exceeded, natural rubbers provide an excellent and economical lining material
to handle fine abrasive and corrosive slurries. Many grades exist, often proprie-
tary. Carbon black, silica, or other fillers and binders are added to increase
strength, hardness, and tear resistance, so as better to withstand the impact of
large particles in the slurry. Rubber is often applied in sheets to large slurry
processing equipment. These sheets may be pre-vulcanized and attached with
adhesives, or applied in the raw state and vulcanized in place. For best results in
high wear applications, such as in cyclones and slurry pumps, rubber liners are
often vulcanized in heavy molds under pressure, requiring specialized
manufacturing equipment. A metal or fiberglass “skeleton” is usually incorpo-
rated into the part to provide strength and allow a point for attachment in the
equipment. This facilitates quick interchange of parts in the field, which is
convenient for maintenance, but limits the flexibility of designers and operators
for frequent design changes, due to the high cost of the molds.
• Neoprene rubbers: Compositions of synthetic neoprene rubber are sometimes
substituted for natural rubber when increased resistance to petroleum products,
higher temperatures, or failure due to tramp material are required. This typically
comes at some sacrifice of wear resistance and increased cost compared to natural
rubber, but these materials fill an important application niche.
• Polyurethanes: Polyurethanes are produced by a chemical reaction between two
or more organic chemicals mixed together in the liquid state and cured into a
highly cross-linked polymer. They are especially effective in applications where
sliding-abrasion wear predominates, and where resistance to tearing and impact is
required, such as screens and launders. They also find wide application in
408 13 Erosive Wear
pipelines, cyclones, and slurry pumps. Many grades can handle a higher flow
energy than rubber, allowing a higher head per impeller stage for centrifugal
pumps. Having less resilience than rubber, their wear resistance is generally lower
under impact wear, but they often exhibit superior performance to both rubber
and hardened metals when the slurry particle sizes are uniformly fine. Their
resistance to temperature is similar to that of rubber, but grades exist with
significantly better resistance to petroleum products. They have the advantage
of a less complex manufacturing process than rubber, being generally mixed into
molds at atmospheric pressure and room temperature, which allows more flexi-
bility for designers and operators. Many different formulations exist, covering a
wide range of resistance to erosion, temperature, and chemical attack, so selec-
tions must be made with care and generally require some testing or experience to
validate performance in slurry applications.
Ceramics
Ceramics comprise the third group of engineering materials commonly applied for
wear resistance in slurry service. These are inorganic compounds of metallic and
non-metallic elements with strong molecular bonds. Ceramics can be very hard and
highly wear-resistant, but are generally low in toughness and impact strength. If we
consider the series of materials that begin with steels and proceed through white
irons, ceramics represent the extremity of the series, the culmination of the trend
towards lower toughness and higher wear resistance. The more successful types of
ceramic materials are also more costly than their metallic counterparts. Applications
in slurry service tend to be limited to non-load-bearing liners strategically placed in
high wear locations such as pipe elbows or chutes, or localized, high wear areas in
centrifugal pumps. They are also used for the rubbing components in mechanical
seals. The most common ceramics in current use for slurry pump liners are
pre-manufactured inserts of silicon carbide and tungsten carbide formed of com-
pressed powders that are molded or machined to shape before being sintered at high
temperature. Alumina (aluminum oxide) is also used, being low in cost and easily
manufactured, but it can often be outperformed by white iron. Some effort has been
made towards producing fully ceramic lined slurry pumps, but to date, these have
found few successful applications. A ceramic’s low tolerance to both shock and
vibration generally requires that, if included as part of a pump assembly, it be
mounted in compression or somehow isolated from the load-carrying components.
Composites
The final category of materials we will consider are composite constructions of
metals, elastomers, and ceramics. A variety of such composites have been applied
with varying degrees of success. These composites are often more complex in their
construction—thus more costly—but have found niche applications in slurry service
wherever they can deliver the best compromise of cost, flexibility, and performance.
Many composites seek to capitalize on the excellent wear resistance of ceramics,
while avoiding their weakness of low-impact strength, by embedding them in a
matrix of epoxy, elastomer, or high alloy metal. A few of these are described below.
13.3 Wear-Resistant Materials 409
Wear in slurry transport applications is caused primarily by the solids, but in addition
to erosion, corrosion can also occur. In former times, corrosion was usually of
significant concern only in chemical or seawater applications, but an increased
focus on strategies to reduce water usage throughout the mining industry has led
to a higher frequency of corrosive content in many slurry flows. It is important to
understand that even small levels of corrosion can influence wear performance
significantly when coupled with the surface effects of particle erosion, since corro-
sion products on the attacked surface layer are usually soft and easily removed by
erosion from the slurry stream (Fig. 13.8).
In a combined corrosion-erosion situation, either erosion or corrosion can be the
predominant factor (Tian and Addie 1996, 1997), and the synergistic effect between
corrosion and erosion can be rather significant and complicated (Tian et al. 2009). In
most slurry transport applications, erosion is the controlling factor, and corrosion
may be difficult to identify as it often occurs together with particle erosion, which
may have a masking effect. Corrosion attack may become much more severe when
acids are found in the slurry, particularly in the presence of impurities such as
chlorides or fluorides. In cases where the suspect chemicals may be separated from
the solids in the slurry flow, erosion and corrosion may be tested separately using a
standardized wear test for comparison with the combined effect on the material. It is
13.5 Cavitation Wear 411
not uncommon for the combined corrosion-erosion effect to be more severe than the
sum of the individual effects. This situation occurs when the particulate erosion
accelerates the corrosion by removing the surface corrosion products, allowing the
fluid access to a fresh surface. In wear testing, samples may be run through
alternating cycles of erosion and corrosion testing to partially simulate this effect.
Where it is suspected that corrosion is a significant factor, and it is not possible to
separate the solids from the corrosive component of the slurry, materials highly
resistant to corrosion (or corrosion-erosion) should be tested as potential replace-
ments for the existing material. The relative particle erosion resistance of these
materials, although important, should not be rigorously imposed to limit the mate-
rials chosen for testing, since the particle erosion component may be greatly reduced
once the corrosion-erosion cycle is broken.
Cavitation may often be controlled through good system design, but in some
applications, such as dredging, it may be in direct competition with production
and cannot be overlooked. Cavitation damage occurs on surfaces where pressure
differences cause the formation and collapse of vapor cavities. When these cavities
collapse near a surface, high-velocity micro-jets of liquid are generated, causing
localized deformation or pitting damage. Cavitation is most often seen near the inlet
of a centrifugal pump impeller, downstream of an obstruction in a pipeline (such as
at a valve or orifice), or anywhere that a sudden drop and recovery of pressure may
occur in a moving flow. The form of cavitation damage is very distinctive and easily
identified when occurring in isolation, having the appearance of cratering or pitting,
or of a severe but localized attack of corrosion.
412 13 Erosive Wear
Wear testing is critical for the prediction and control of wear in slurry handling
equipment such as pipelines and pumps. Numerous studies have been conducted to
quantify and characterize wear in a system, including those of Finnie (1960), Tuzson
et al. (1984), Miller (1987), Roco et al. (1987), Pagalthivarthi and Helmly (1990),
Sundararajan (1991), Hutchings (1992), Tian et al. (2005, 2009), Li (2005), Laguna-
Camacho et al. (2013), Xie et al. (2015), and Blau (2017). In any wear test, the goal
is to rank either the wear resistance of several materials or the abrasivity of several
solids, or to obtain an absolute wear coefficient (energy consumed per unit volume of
material removed) for some specified combination of the two. The latter is typically
used for predicting wear rate and effective life.
In selecting a wear test, attention must be paid to the types of wear to be
encountered under the actual operating conditions, and how well these can be
simulated. For instance, the operating environment might include long-term corro-
sive damage, which is difficult to implement experimentally. Tests must also provide
realistic and measurable wear rates. Often these requirements are contradictory,
since actual wear rates may be too small to permit accurate measurement within
the period of testing. Conversely, tests with artificially increased wear rates may not
represent prototype wear adequately, so that the scale-up of erosion measurements is
complicated.
In planning a wear test, three factors must be taken into account: slurry charac-
teristics, material characteristics, and operating characteristics of the equipment. The
wear characteristic of the slurry, also known as slurry abrasivity, will be determined
by the particle size distribution and particle shape (whether angular or rounded),
hardness, and friability. Large particles tend to cause impact damage, and angular
particles are instrumental in scratching and scouring the surface. The hardness of the
particles in relation to that of the surface material is very important in determining
crack propagation, especially when impact damage is of concern. Also, hard
13.6 Experimental Testing Methods 413
particles tend to retain their sharp edges after scratching the surface but are likely to
be friable upon impact. For this reason, they may lose some of their ability to cause
further damage, although the new sharp edges resulting from particle breakage can
produce additional scratching. The erosive ability of the slurry is compounded if it
has a non-neutral pH or other corrosion-inducing factors such as halogen elements or
ions; in this case, the added mechanism of corrosion will be significant.
The most important material characteristics that affect wear are the chemical
composition, hardness, microstructure, toughness, and surface finish. If the material
is harder than the particles, the particles will be unable to scratch the wear surface
effectively. Toughness of the material is significant when the wear mechanism is
fatigue controlled, and for many applications a good combination of hardness and
toughness is necessary. Important microstructural considerations are the distribution
of second-phase materials such as carbides in iron alloys, and also the grain size and
the orientation and shape of the second phase. The chemical composition of the
material is especially significant in corrosive media. Finally, the surface finish can
have a significant effect on crack propagation.
A detailed description of the various types of test apparatus used for quantifying
wear and ranking wear resistance of materials cannot be undertaken here, but salient
features of some of the most relevant tests deserve mention.
The Miller Machine
The Miller machine is commonly used to rank both slurries and wear-resistant
materials in the order of their abrasive wear potential or resistance (Miller 1987).
In this test, a wear sample is made to reciprocate for a specified number of cycles in a
slurry of typically 50% by weight concentration of solids. A fixed weight is used to
impose a normal force between the slurry and the sample, and the sample face is
chamfered to force the slurry mixture between it and a stationary rubber lap.
Measurements of the rate at which the sample loses mass serve to determine either
the slurry’s abrasivity or the material’s wear resistance. In the former case, a
standardized, centrifugally cast white iron sample is used against the slurry of
interest, and a Miller number is calculated which indicates the slurry’s abrasivity,
which can be used for ranking the wear potential of the slurry (see Fig. 14.2). In the
latter case, a standardized mixture of silica sand is used against the wear material of
interest and a Slurry Abrasion Resistance (SAR) number for the material is deter-
mined. (Note that other types of wear test rigs can also be employed for this two-way
use, measuring either wear resistance of materials or abrasivity of slurries.)
The Miller test is both easy to perform and widely known, making it useful for
comparative ranking of slurries and materials. It is most useful for applications
where heavy sliding-abrasion dominates. However, because it is a three-body wear
test where the slurry particles are forced between the wear sample and a rubber lap, it
is less representative of the two-body slurry erosion seen in many types of processing
equipment, where a free slurry flow runs against a fixed wear surface. It can provide
only relative wear resistances or abrasivities between various materials or slurries
and cannot provide the absolute wear coefficients required to estimate wear rates
under a given set of operating conditions in a pipeline or pump.
414 13 Erosive Wear
Fig. 13.9 Coriolis wear machine. End cover rotated and sample extended at left
Testing at GIW has shown that the exact pattern of wear on identical specimens can
vary considerably while still maintaining repeatability in overall weight loss and
allowing correlation to the energy expended in creating the wear. This indicates that
the random or chaotic nature of wear parallels that of fluid turbulence, where
significant real-time fluctuations are often superimposed on an apparently steady
(time-averaged) flow.
Flow within the Coriolis machine is also easily modeled, either analytically as
done by Tuzson et al. (1984), or numerically as done by Pagalthivarthi et al. (2017a).
Both approaches allow for calculation of the energy dissipation at the wear specimen
surface, which, combined with the experimental wear rate, can be used to determine
the absolute sliding wear coefficients. These coefficients relate the amount of
material removed to the energy dissipated (Tian et al. 2005). Wear coefficients are
essential inputs to the numerical simulation of slurry equipment wear, as described in
Sect. 13.7.
Impact Wedge Test
The wedge test shown in Fig. 13.10 is useful for determining impact wear coeffi-
cients. In this test, wear specimens with various angles on their leading edges are
inserted into a rectangular channel through which a standardized silica sand slurry is
pumped. In theory, the particle incidence angle, impact velocity, and concentration
can be determined by the inlet angle of the specimen, along with the average velocity
and concentration of the slurry flow. In practice, all three quantities will vary along
the specimen, due to the diversion of the flow caused by the specimen itself. This is a
problem with virtually all slurry flow impact wear experiments.
In the wedge test, the geometric impact angle with average velocity and concen-
tration will be most closely represented at the tip of the specimen, so one can make a
series of careful measurements of actual wear depth at evenly spaced lines down-
stream of the tip and extrapolate these values back to the tip (often worn away) to
obtain useful experimental results. Better information can be obtained by performing
416 13 Erosive Wear
called the wear coefficient (Wc). When modeling the slurry equipment of interest,
surface dissipation energy at any point is calculated by numerical simulation
(or other means) and the wear rate at that point is predicted by the application of
the wear coefficient.
It must be understood that each wear coefficient is tied to a particular combination
of slurry type, wear material type, and erosive wear mechanism. Consequently, there
may be many different wear coefficients for different uses. To simplify the com-
plexity, a set of standard coefficients is often generated for a common combination
of slurry, wear material, and erosive mechanism (for example, silica sand against
high chrome white iron in sliding-abrasion wear). This coefficient can then be
applied to a wide range of other slurries and materials using experience-based
multipliers, as long as the basic mechanisms of wear do not vary too greatly.
Thus, a wear coefficient based on silica sand against high chrome white iron might
easily be applied to copper ore against chrome-moly white iron using appropriate
multipliers, but separate wear tests and a new coefficient may be needed if the wear
material were an elastomer.
For wear modeling of centrifugal slurry pumps and pipelines, two main erosive
mechanisms predominate: sliding-abrasion and impact wear. As a result, a wear
coefficient for each is required. The sliding-abrasion wear coefficient will usually
vary with particle size, while the impact wear coefficient will also vary with impact
angle. To obtain these dependencies, a range of wear tests will be required for the
slurry and material type of interest. Figures 13.11 and 13.12 show typical sliding-
abrasion and impact wear coefficients for silica sand slurry against a variety of
common wear materials. Note the variations in particle size dependency for sliding-
900
800
Hi-Cr White Irons
Typical Sliding Wear Coefficient
700
Hyper-Eutectic
WC x 10E-14 (m3/J)
600
White Irons
500
Elastomers
400
300 Ceramics
200
100
0
0 1000 2000 3000 4000 5000 6000
Particle Size, d (µm)
Fig. 13.11 Typical sliding wear coefficients for silica sand against various common wear materials
in a neutral pH medium as a function of particle size
418 13 Erosive Wear
Ductile metal
(e.g. steel)
Relative impact erosion rate
per unit volume of solids
White Iron
Elastomers
(e.g. 50A rubber)
Ceramics
0 30 60 90
Impact angle (degrees)
Fig. 13.12 Impact wear coefficient trends for typical slurry pump wear materials. The curves
display the form only and are normalized with respect to their maximum values. In practice, the
magnitudes will vary with material grades and slurry conditions
abrasion wear displayed by the different material types. These can be seen to
correspond to the above discussion on wear materials. In a similar way, the impact
wear coefficients show a strong dependence on the impact angle, which also varies
greatly with the type of wear material. The hardest materials tend to show their
highest impact wear rates at high angles of impact, where energy transfer is maxi-
mized; the more ductile materials display a peak wear rate at some intermediate
angle where cutting and/or deformation wear are maximized.
The availability of high-speed computers with ample memory has encouraged many
researchers to develop numerical algorithms for analyzing flow and attendant wear
in pumps and other components of slurry pipeline systems. Since wear depends on
local values of velocity and concentration, the flow and concentration fields must be
computed prior to wear evaluation. Either an Eulerian-Lagrangian or an Eulerian-
Eulerian approach is generally used to compute the flow and concentration fields in
the wetted components of pipeline systems.
In the Eulerian-Lagrangian approach, the fluid phase is treated as a continuous
medium, and the relative motion of the discrete solid particles is individually tracked
with respect to the fluid flow. This approach is most suitable for very lean slurries
with relatively larger particles, implying relatively fewer particles to be tracked and
13.8 Numerical Modeling of Flow and Wear 419
bulk-fluid solution. This process is repeated until the solution converges on a stable
result (Fig. 13.13).
A two-dimensional approach is less meaningful for analyzing the impeller, due to
the complex geometry and the lack of a clear plane for analysis which represents the
action of the entire flow field. The secondary flows (large-scale eddies at right angles
to the main flow) represent a significant part of the fluid motion due to the oversized
passages typical of slurry impeller designs meant to handle larger solids. These
secondary flows cannot be ignored, and their computation is more complex, as the
viscous forces are of the same order of magnitude as the inertial forces and must be
included in constructing the governing equation of the numerical model. The same
applies to three-dimensional analysis of the casing, especially at flow rates away
from the best efficiency flow rate (QBEP), where secondary flows increase
(Visintainer et al. 2019). Secondary flows can vary significantly depending on the
operating conditions, the pump geometry, and the assumptions made about the inlet
boundary conditions. The complex outlet velocity profile of the impeller represents
the inlet boundary condition for the casing model, and this profile has a particularly
strong influence on performance. Work done by Hergt et al. (1994) has shown that
typical impeller outlet profiles in slurry pumps are quite unlike those encountered in
conventional water pump designs. An example of secondary flow patterns in a slurry
pump impeller and casing is shown in Fig.13.14.
Early modeling attempts by Roco and Reinhardt (1980), Roco and Addie (1983),
and Roco et al. (1984a) began by calculating the inviscid flow field at the compu-
tational grid points. The particle velocity was then computed at each node using a
force balance, taking into account particle inertia, centrifugal force, and fluid resis-
tance. If the solids concentration is small, the consideration of wear could begin at
422 13 Erosive Wear
this point in the analysis, with sliding-abrasion occurring where the particles move
along the boundary, and particle impact erosion where the trajectories arrive at an
angle.
For conditions typical of slurry transport, however, near-wall concentrations are
high. In these circumstances, the question of wear can only be addressed after the
fraction of particles supported by granular contact is known. The process of deter-
mining the fraction of contact-load solids in a pump casing is basically similar to that
of evaluating the stratification ratio for heterogeneous pipeline flow, a problem that
was discussed extensively in Chaps. 4, 5, and 6. As noted in those chapters, even for
the simple configuration of pipeline flow, the physical mechanisms determining
stratification are complex, and achieving better understanding remains a major
research aim. For the more complicated configurations encountered in pumps,
there is even less certainty as to the action of the various physical mechanisms
involved. An approach to this problem which has received considerable attention in
the literature is due to Roco and his co-workers. This researcher worked at the
University of Paderborn in Germany (Roco and Reinhardt 1980) and subsequently
was associated with Shook in Saskatchewan (Roco and Shook 1983, and numerous
other joint publications). During the mid-1980s, Roco worked in conjunction with
the GIW Hydraulic Laboratory (Roco et al. 1984a, b, 1987), where the modeling of
wear in pumps remained an ongoing priority (Pagalthivarthi and Addie 1989;
Pagalthivarthi et al. 1990, 1991).
After the fraction of contact-load solids has been estimated at various locations
within the pump, the procedure continues to the remaining step: the consideration of
erosion rates. In early algorithms, the local erosion rate was taken as proportional to
the rate at which energy is lost to friction and other dissipative mechanisms, and the
local energy flux of the particles at the wall was related through empirically
determined coefficients to the predicted wear rate.
13.8 Numerical Modeling of Flow and Wear 423
The total wear rate may be computed as the sum of an impact component and a
sliding component, however, this linear superposition may not be strictly valid.
Some researchers (e.g., Roudnev and Kosmicki 2009) employ an overall wear
coefficient to circumvent breaking the total wear into its component parts. Deter-
mining the overall wear coefficient would involve conducting expensive wear tests
on actual pump casings (Pagalthivarthi and Helmly 1990) and impellers for different
particle sizes, abrasivity, and perhaps even concentration. Such tests typically take
much longer durations to yield measurable wear. With such extended-duration tests,
the solids must be repeatedly replaced with fresh batches to counteract the break-
down of particles due to circulation through the test rig. On the other hand, in the
Coriolis and wedge tests, measurable wear is seen in just an hour or two of operation,
even with relatively wear-resistant materials, and it is possible to conduct multiple
tests with several wear materials, particle types, and shapes in a matter of days. Thus,
from a pragmatic consideration, there is a trade-off between using an overall wear
coefficient versus splitting it into sliding and impact components.
The current state-of-the-art modeling of dense slurry flow involves a Eulerian-
Eulerian approach (Pagalthivarthi and Visintainer 2013a, b; Pagalthivarthi et al.
2015), where both the carrier fluid and the solids are treated as continuous media.
A streamlined upwind Petrov Galerkin–based finite element formulation is used to
model the flow inside the casing and impeller. Simulation is performed in two stages.
In the first stage, the carrier fluid phase is solved using a one-equation Spalart-
Allmaras turbulence model (Spalart and Allmaras 1994). This is done primarily to
obtain the eddy viscosity distribution in the flow domain for the carrier fluid flow. In
the second stage, the slurry mixture and solids equation (velocity and concentration)
are solved, and the eddy viscosity obtained from the carrier fluid phase is modified to
include the effect of concentration. Penalty formulation is used for the carrier phase
and the mixture velocity (momentum) equation. Spalding wall functions (White
1991) are applied at the wall, and a parallel multi-threaded sparse direct solver, such
as Intel MKL PARDISO (Schenk and Gartner 2004) is used to solve the linear
system of equations. The results from this CFD procedure have been compared with
experimental flow field data within a centrifugal casing (Furlan et al. 2015) and a
rectangular three-dimensional duct (Pagalthivarthi and Visintainer 2013a).
When modeling slurry flows, the particle size distribution (PSD) may be
represented by one “effective”, mono-sized particle (Pagalthivarthi and Visintainer
2009), or divided up into multiple fractions or “species” which are treated separately
for improving the accuracy of the wear prediction (Pagalthivarthi and Visintainer
2013a, b; Pagalthivarthi et al. 2015). Pagalthivarthi et al. (2017b) compared mono-
sized and multi-species models and found that an average of the d50 and d85 particle
sizes provided a reasonable representative size for the mono-sized analyses, provid-
ing similar wear results to the multi-species analyses, provided that the PSD was of a
normal form. Since mono-sized particle simulations are significantly faster than
multi-sized simulations, this provides a convenient time-saving approach for routine
analyses.
The wear rates along the wall surfaces are calculated using the local flow
quantities near the wall and the wear coefficients estimated from the Coriolis
424 13 Erosive Wear
(sliding) and impact wedge wear tests (Pagalthivarthi et al. 2017a and Sect. 13.6).
The total wear is then calculated as the sum of the sliding and impact wear
contributions.
The impact wear rate of an individual particle size is expressed as a function of the
kinetic energy flux of those particles. With the total rate given as a summation of the
individual contribution of each particle species.
X
N
ρk ck u3k
_I=
W ð13:1Þ
k=1
EI ðαk Þ Cadj
where ρk, ck, uk, and αk are, respectively, the density, concentration, velocity and
impact angle of particle species ‘k’, and EI is the specific energy coefficient
(or impact wear coefficient) for the kth species. The term ρk ck u3k is the impact flux.
Based on experiments, the particle size dependence of EI (J/m3) may be written as:
where EI(αk, d0) is the value for the reference particle size d0, and Cadj is an empirical
factor adjusting for the actual particle size (dk). Typical formulations are:
1
EI ðαk , d0 Þ = cos n ðαk Þ
ð13:3Þ
þ sin Bðαk Þ
n
A
n
C adj = Aim dpk þ C im im þ Bim ð13:4Þ
with A, B, n, Aim, Bim, Cim and nim being determined experimentally. The exponent
n varies with α between 2.5 and 3.0, and A and B represent the estimates of the
impact wear coefficient for 0° and 90° impact angles. With dk expressed in microns,
αkin radians, and EI in (J/m3), a high-quality chrome white iron may be modeled
using the following constants:
d 0 = 270 μm,
A = 1:5 × 1015 ,
B = 3 × 1011 ,
Aim = 6:19 × 1010 ,
Bim = 3:645 × 10 - 2 ,
C im = 390,
nim = - 3:668
XN
ck τk utk
_ S=
W ð13:5Þ
k=1
E sp ðd k Þ
where ck is the local concentration near the wall, τk is the shear stress caused by the
particles on the wall, and the utk is the velocity of the particles tangential to the wall
surface. The sliding wear coefficient (in J/m3) is a function of the particle diameter,
given as:
where Asl, Bsl, Csl and nsl are empirically determined constants for a particular
material. With dk expressed in microns, and ESP in (J/m3), a high-quality chrome
white iron may be modeled using the following constants:
For consistent wear prediction, the empirical constants in the wear coefficients are
obtained from Coriolis and impact wedge tests where the local wear measurements
are correlated to local flow quantities at the wall determined using CFD simulations
of similar formulation to those used in modeling the slurry pumps themselves
(Pagalthivarthi et al. 2017a).
As mentioned previously, numerical simulation of turbulent flows is complex and
subject to many potential errors, due to the nature of the discretized computing
process and the assumptions made to simplify the governing equation of motion.
Therefore, for successful modeling, these simulations must be refined and validated
by experimental measurements. Ideally, these should match, as closely as possible,
the slurry properties, flow conditions, and geometry being modeled. Much, however,
can be achieved with simplified geometries and flow fields, provided they display
similar hydraulic mechanisms. Without such experiments against which to test and
refine these models, the degree of error introduced by the computational inaccuracies
and mathematical assumptions can lead to substantial errors.
In experimental work with slurry flows, the measurement of solids concentration
and velocity is often complicated by the opaque nature of the slurry and the erosive
action of the particles (which may damage the experimental equipment). Specialized
techniques may be required, such as the use of artificial slurries composed of clear
solids and liquids with similar indexes of refraction, or the use of sophisticated
ultrasound techniques to “look” within the slurry and determine its properties over a
wide region. Notable work accomplished in this field by Kadambi et al. (2004) and
Furlan et al. (2009, 2012, 2015) has provided essential data for the refinement and
validation of numerical methods for wear prediction in slurry pumps.
Of particular interest was the experimental measurement of the solids concentra-
tions and velocities at the outlet of a typical impeller (Furlan 2011; Garman 2015).
For a steady flow of fluid and solids through a centrifugal slurry pump, the mea-
surements showed an average in-situ volumetric concentration of solids at the
impeller outlet that was lower than the average value at the suction inlet. Since
426 13 Erosive Wear
mass continuity requires that the delivered concentration be identical at both loca-
tions, the different in-situ concentrations had to be a result of different slips between
particles and liquid. (See Sect. 3.2.2 for the effect of slip on concentration.) Although
perhaps counter-intuitive, the result was confirmed by multiple experiments, and,
upon reflection, it makes physical sense, as the higher density particles experience
higher centrifugal forces within the rotating flow field and are thus accelerated to
higher velocities than the surrounding liquid. This produces “negative” slip and
results in a lower volumetric concentration. The process is reversed in the casing,
where the particles are preferentially decelerated by friction against the casing wall,
producing “positive” slip where the particles move slower than the surrounding
liquid. Due to the conditions in the casing, the concentration of particles at the casing
outlet becomes higher than at the impeller inlet.
Another less precise (but equally useful) validation method comes from compar-
ing the results of numerical methods to experimental measurements taken from
actual production equipment in the field. While such field experiments are usually
less controlled than their laboratory counterparts, both in operating conditions and
measurement accuracy, the sheer volume of such opportunities creates a large body
of experimental evidence that, over time, helps to calibrate and refine the use of
numerical models. In many cases, the models are used to examine improvements to
an existing installation and, as such, can be calibrated to the wear performance of the
existing equipment before potential modifications are analyzed.
Although various remarks have been made throughout this section concerning the
limitations of numerical wear analysis, this technique remains of significant value to
the design engineer. The results of numerical analysis are especially useful when
comparing alternative designs, and the capability of predicting local wear rates can
be very helpful in avoiding catastrophic erosion failures. For these reasons, the
numerical and experimental study of liquid-solid flows has been particularly empha-
sized at the GIW Hydraulic Laboratory.
As noted previously, much wear performance is evaluated post facto. This approach
is necessarily limited when it comes to predicting trends over a wide range of
operating conditions that may span gaps in the available experiential data or extrap-
olate beyond them. Numerical modeling, as described above, can provide a useful
tool for revealing these trends. Although many variables contribute to pump wear
performance—some related to slurry properties, others to pump geometry, and still
others to the interactions between pump and system—much understanding can be
gained by performing ordered numerical experiments over a range of key parameters
(Sellgren et al. 2005; Addie et al. 2005).
One such parametric study by Visintainer et al. (2022) considered a collection of
112 different slurry pumps selected to represent the range of designs in current use.
These included impeller vane diameters (D) from 125 to 2667 mm and pump-
13.9 Parametric Study of Slurry Pump Wear 427
specific speeds (Ns) from 11 to 105. Each design was numerically modeled over a
range of operating conditions, including variations in head (H ), flow rate (Q),
delivered volumetric concentration (Cvd), solids density (ρs), and effective solids
size for wear (deff), which was taken as the average of the d50 and d85 from the
particle size distribution. Each operating condition was varied relative to a baseline
case of H = 50 m, Cvd = 0.2, ρs = 2650 kg/m3, deff = 500μm, and Q/QBEP = 1.0,
where QBEP is the best efficiency flow rate at a given head. In the resulting analysis,
the effects of variable operating conditions were considered in tandem with the
effects of pump size, specific speed, and other representative geometric ratios of the
designs.
Other important variables affecting wear are the abrasivity of the solids and the
wear resistance of the materials. These variables are experimentally correlated to the
energy required to remove a unit volume of material and incorporated into wear
coefficients (see Sect. 13.7), which then become input variables to the numerical
wear models. For the purpose of the study, wear coefficients representative of semi-
angular silica sand against a high-quality 28% chrome white iron were used. In
practice, results for one abrasivity or material resistance are scalable to other cases
(given the same operating conditions), according to the ratio of the wear coefficients.
Fluid density and viscosity can also affect wear; however, the conveying fluid is
water in many cases of industrial interest, so for the purpose of the study, these
properties were held constant at 1000 kg/m3 and 1.0 cP, respectively.
The numerical models used included two-phase, finite element CFD models of
slurry pump flow and wear as described above—in particular, a 2D casing wear
model (Pagalthivarthi and Visintainer 2013b), a quasi-2D suction liner wear model
(Addie and Bross 2000; Bross and Addie 2001), a 3D impeller wear model
(Pagalthivarthi et al. 2013a), and a 3D casing wear model (Pagalthivarthi et al.
2013b, 2015). Development of these models over time has advanced in tandem with
increases in available computational power, allowing an ever-increasing complexity
in the constitutive equation of flow and solids-fluid interaction used in the models.
This has included the introduction of increasingly complex formulations for han-
dling the viscous terms in the Navier-Stokes equation and in modeling the stochastic
effects of turbulence.
In developing correlations of the modeled results, the selected operating param-
eters were nondimensionalized by division with some baseline value, and most of
the geometric variables were also represented by nondimensional ratios. In a few
cases, dimensional values were used, and these were especially important in captur-
ing the effects of pump size. In general, a correlation equation of the following form
was used:
e1 e2 e3 e4 h ie5
_ = C0
W H deff Cvd ρS
1 þ abs QQBEP - Q0 ϕe6
6 ϕ7 ϕ8 . . .
e7 e8
H0 d0 C V0 ρS0
ð13:7Þ
428 13 Erosive Wear
where W _ is the wear rate (in μm/h) and the variables H0, d0, CV0, ρS0, and Q0
represent baseline values for each of their associated quantities. The variables ϕ6,
ϕ 7, ϕ 8, . . . represent additional selected parameters, usually geometric in nature;
the variables e1, e2, e3, . . . represent the power to which each parameter is raised;
and the variable C0 is a fitting constant.
By their nature, CFD models are approximations, and simplifications must be
made in any parametric analysis; therefore, care must be exercised in applying these
correlations to real-life conditions. While they can be used for “order-of-magnitude”
estimations of pump wear given an arbitrary set of inputs, they are most useful for
the trends and dependencies they illustrate, and they are especially helpful in
predicting the expected incremental changes in wear performance resulting from
an incremental change to one or more parameters relative to an established baseline
condition, for which experience already exists.
The controlling wear rate for a suction liner is generally at the clearance gap between
the stationary liner and the rotating impeller. The strongest correlating variable for
geometry in the predicted suction liner wear was the ratio between the impeller vane
diameter (D) and the impeller suction eye diameter (De), as defined in Fig. 13.17.
Figure 13.15 shows results for the baseline operating condition and a typical suction-
side clearing vane geometry. Wear is inversely proportional to the ratio (D/De),
following a power curve to the exponent of -3.25.
100
Suction Liner Wear Rate (Pm/hr)
10
wear ~ (D2/De)-3.25
1
1.0 2.0 3.0 4.0 5.0 6.0
Impeller Vane Diameter D2 / Suction Eye Diameter De
Fig. 13.15 Suction liner wear rate for the baseline operating condition. (Visintainer et al. 2022)
13.9 Parametric Study of Slurry Pump Wear 429
Considering all of the results for variable operating conditions, the following
_ SL ), in μm/h, was developed:
generalized correlation for suction liner wear rate (W
Note that the clearing vanes, which are located on the outside of the suction-side
impeller shroud, can have a significant effect on the flow in the gap between the
impeller and side liner, and thereby on the liner wear rate. Additional terms for their
effect were included in the original work.
In evaluating casing wear, the maximum wear rate is usually considered at any given
point since it determines the life of the part. If operated in normal ranges of
(Q/QBEP), casings will usually wear out at the periphery, or the “belly,” as it is
sometimes called. The strongest correlating variable for geometry in predicted
casing wear was related to the pump power factor and the centripetal force of the
particles in the casing domain. It is a dimensional (size-affected) parameter equal to
the volume of the casing outside the impeller (VCAS), multiplied by the average
radius of the casing passages (rA). Figure 13.16 shows this parameter plotted against
the maximum calculated 2D casing wear rate at the baseline operating condition.
100
Max. casing centerline wear rate (Pm/hr)
10
1
0.000 0.001 0.010 0.100 1.000 10.000
VCAS · rA (m4)
Fig. 13.16 Maximum casing wear rate for the baseline operating condition. (Visintainer et al.
2022)
430 13 Erosive Wear
Fig. 13.17 Geometric variables associated with impeller and casing wear rate correlations.
(Visintainer et al. 2022)
Note that the wear rate is inversely proportional to (VCAS rA), meaning that larger
casings will wear somewhat less than smaller ones, with the power of proportionality
being approximately -0.25.
The parameter (VCAS rA) may be estimated from a few key casing dimensions
using the following equation:
where (VCAS rA) has units of m4. Note that the dependencies of the operating
parameters of head, particle size, density, and concentration differ from those found
for the suction liner due to the different nature of the flow field. In the case of the
suction liner, the flow field is highly rotational, while in the casing, wall friction
plays an important part in developing the concentration distribution and solids
velocities which control the wear rates. The dependence on (Q/QBEP) is stronger
and finds a minimum near the lower flow rate of (Q/QBEP) = 0.75. Of the remaining
13.9 Parametric Study of Slurry Pump Wear 431
Determining the controlling wear rate for impeller life is more complicated than for
the casing or suction liner. For those parts, the maximum wear rate determines the
point at which the part wears through and begins to leak. The impeller, on the other
hand, can sustain considerable wear in a localized area while still maintaining
acceptable performance (see Fig. 13.1). Using a 3D impeller wear model, we can
examine different zones of the impeller and, combined with experience, make
reasoned choices regarding the controlling wear rate. For a parametric study, how-
ever, we must choose a rationalized value, such as the average wear rate over the
entire surface of the vane. While this is simplified, average vane surface wear does
represent the region of most concern with regard to performance and will often be the
controlling factor for wear life, except in cases where very large solids are handled
(Visintainer and Wolfe 2015).
The strongest correlating variables for geometry in the predicted impeller wear
were the pump-specific speed Ns, flow rate Q (m3/s), impeller vane diameter D (m),
and total vane area AV (m2), which includes the pressure and suction sides of all
vanes combined. Figure 13.18 shows the results for the average calculated vane wear
against these four parameters at the baseline operating condition.
100
wear ~ (Ns·Q)/(D2·AV)0.7
Average vane wear rate (Pm/hr)
10
1
10 20 50 100 200
( Ns · Q ) / (D2 · AV)
Fig. 13.18 Average impeller vane wear rate for the baseline operating condition. (Visintainer et al.
2022)
432 13 Erosive Wear
The wear rate is directly proportional to the specific speed and flow rate and is
inversely proportional to the impeller diameter and total vane area. Considering all
of the results for variable operating conditions, the following generalized correlation
for the average impeller vane wear rate W _ IMP , in μm/h, was developed:
where the units for Q, D, and AV as noted above are used. This wear rate includes
both the suction- and pressure-side vane surface wear rates, which are added together
at each point on the vane, and then averaged over the entire vane area.
As with casing wear, an inverse size effect with impeller wear becomes visible in
the D and AV terms. The minimum wear occurs near (Q/QBEP) = 0.85. The effect of
head is not as strong as in the suction liner or casing correlations, perhaps because
the impeller is rotating with the flow field, thereby reducing the relative velocities at
the wear surfaces normally associated with increased head.
A formulation of similar accuracy, but one that does not require the flow rate or
impeller vane area, is as follows:
where Z = the number of vanes. In this formulation, the effects of flow rate
magnitude and vane area are “bundled” into the QBEP correction, vane number,
and pump-specific speed terms. While the first equation is preferred as being closer
to the physical principles, the second equation is more convenient for everyday use.
By experience, wear often has some dependency on pump-specific speed and size.
These factors may not provide the best correlating variables in isolation, but because
they are global variables encompassing a wide range of effects, it is worthwhile to
examine their relationship to pump wear. In Fig. 13.19, the controlling wear rates
(as described above) are plotted against pump-specific speed (Ns). There is a clear
upward trend in suction liner and impeller wear with specific speed, while that for
casing wear is weaker and downward trending. The difference between the 2D and
3D casing wear models is also shown, with the 3D results having a slightly stronger
dependence on specific speed.
13.9 Parametric Study of Slurry Pump Wear 433
100
Controlling Wear Rate (Pm/hr)
10
1
0 10 20 30 40 50 60 70 80
Pump Specific Speed Ns
Fig. 13.19 Wear rate as a function of specific speed for the baseline operating condition.
(Visintainer et al. 2022)
100
Controlling Wear Rate (Pm/hr)
10
1
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Impeller Diameter D (m)
Fig. 13.20 Wear rate as a function of vane diameter for the baseline operating condition.
(Visintainer et al. 2022)
Figure 13.20 shows the controlling wear rates plotted against the impeller vane
diameter (D). An inverse trend is seen for casings and impellers, indicating that
larger pumps wear at lower rates, a result confirmed by field experience. The suction
liner trend is slightly increasing, but the correlation is poor and probably influenced
434 13 Erosive Wear
10,000
Wear Life (hr)
1,000
100
0 10 20 30 40 50 60 70 80 90 100
D (m) · Ns
Fig. 13.21 Predicted parts life for the baseline operating condition. (Visintainer et al. 2022)
by there being more high-specific speed pumps at larger sizes in the data set. The 3D
casing results show a slightly weaker dependence on vane diameter relative to the
2D.
There is a significant scatter in both plots because each omits the effect of the
other, as well as other geometric variables identified as important in the previous
discussion. Some improvement in the correlation can be made by combining specific
speed and size into one factor (D Ns), as seen in Fig. 13.21. In this figure, the wear
rates have been converted to wear life based on standardized part thicknesses. A
divergence in part lifetimes, with suction liner wear becoming especially low, is seen
with an increasing value of the (D Ns) factor. An approximate equivalence of parts
life is seen in the region of (D Ns) = 20 to 30, which indicates a desirable design
target for heavy-duty slurry service, especially for small to mid-sized pumps (i.e.,
D < 1 m). For larger pumps, achieving a low enough specific speed to reach the
equivalence point is usually not economical, and some differential is usually
expected, with a ratio of two or three suction liners and two impellers for each
casing being common.
In practice, operating duties vary, and slurry pumps do not always operate near the
optimal points on the curves for wear and efficiency. Pump designs vary a great deal,
and as might be expected, wear varies considerably from pump to pump, and from
13.10 Practical Considerations and Field Experience 435
Near Volute
Semivolute
Recirculation within the casing may also affect flow patterns in the clearance
between the impeller and the suction-side liner, producing secondary gouging there.
The localized wear can be severe (see Fig. 13.25). In such cases, a practical method
of extending the life of the liner is to rotate it 180° when one-half to three-quarters of
its estimated life has elapsed, thus moving the wear zone to a new portion of the
surface. However, resizing the wet end in order to place the pump QBEP closer to the
operating condition is usually a more effective solution for eliminating gouging wear
in both the liner and casing. Controlling the impeller to suction-side liner clearance
can also help to reduce gouging wear, as described further below.
Gouging eddies may occur at other locations in the slurry pump casing or in the
impeller. In some instances, eddies can be identified as having their origin in
geometric discontinuities or poor hydraulic design, and they can be eliminated by
judicial design modifications. However, in many cases, they may be traced to
operational considerations, with operation well away from QBEP being the most
common. In all cases, it is vital to have a sound understanding of the large-scale flow
patterns within the pump. This understanding depends on careful observation and
also on analysis, using both basic fluid mechanics and numerical modeling.
Another important consideration for controlling slurry pump wear is to control
the clearance between the rotating impeller and the stationary suction-side liner,
sometimes referred to as the nose gap (Fig. 13.26). Some fluid will always
recirculate through this clearance from the higher-pressure casing to the lower-
pressure suction inlet. In a slurry pump, this flow carries particles that cause wear,
438 13 Erosive Wear
NOSE
GAP
and as the clearance opens, recirculation and wear will increase. In most slurry
pumps, this is an axial (perpendicular to the shaft centerline) rather than a radial
clearance, as seen in many clear liquid pumps. A radial clearance has the advantage
of being automatically set when the pump is assembled, but when particles are
present, the centrifugal forces within the fluid of the gap cause them to run against
the outer ring, which quickly wears out. The axial clearance used in most slurry
pumps avoids this problem, and it can also be re-adjusted after wear by axial
movement of the impeller or suction-side liner. Controlling this clearance over the
life of the pump by frequent adjustment will usually increase the life of both the liner
and impeller by 25–75%, and this practice has become standard for many slurry
pumps in heavy-duty applications. For optimal performance, the adjustment should
be made 12–25 times during the life of the parts—for example, once or twice a week
for a pump impeller and liner that are expected to last for 3 months.
Impeller wear is often closely related to the hydraulic efficiency of the impeller’s
design, all other things being equal. This seems reasonable since improved hydraulic
efficiency generally coincides with reduced energy dissipation and reduced veloci-
ties in the recirculating eddies that cause localized wear. However, a well-designed
slurry pump impeller can often sustain considerable wear before its pumping
capacity is reduced to an unacceptable level, and it may appear to be worn out
long before it needs to be replaced from an operational standpoint (see Fig. 13.1).
Wear at the impeller inlet is often the result of recirculation caused by operation at
low flow (well below QBEP). This wear tends to destroy the sealing face of the
13.10 Practical Considerations and Field Experience 439
Fig. 13.27 Inlet recirculation wear of an impeller caused by low flow operation
impeller, increasing recirculation and wear even further (Fig. 13.27). High flow
(high inlet velocity) can also be a problem for slurry pump impellers, as these can
cause the heavier solids in the flow to impinge on the back shroud of the impeller,
wearing through. If left unchecked, this problem can completely separate the
impeller from its shaft connecting hub (Fig. 13.28).
The wear of other components of a slurry pipeline transport system is also
important. The pipe itself can wear, of course, as a result of both particle impact
associated with fluid turbulence, and sliding-abrasion. The latter mechanism is
usually dominant in pipe flow, with wear concentrated in the lower portion of the
pipe for both fully stratified and heterogeneous flow regimes. Sometimes the
resulting wear can limit pipe life; in this case, it may be worthwhile to rotate the
pipe through 120° at the estimated one-third and two-thirds points of its life span,
thus limiting the wear at any one part of the pipe wall. In more serious cases, a lined
pipe may be considered. A stationary deposit of solids in the pipe produces a
characteristic wear pattern, with the wear concentrated in bands at each side of the
lower portion of the pipe interior, say at the 4 o’clock and 8 o’clock positions.
Sometimes the occurrence of this wear pattern provides the first indication that a
system has been operated at too low a mixture velocity (i.e., in the stationary-deposit
flow regime).
In bends and other fittings, wear is usually much more severe than in straight
lengths of pipe, with wear rates easily increasing by an order of magnitude or more.
The basic reason for this enhanced erosion is the curvature of the flow stream, which
produces centrifugal forces that cause the particles to impact the boundary, a point
noted earlier in this chapter. Use of a larger bend radius to reduce these forces is a
440 13 Erosive Wear
Fig. 13.28 Impeller back shroud “blow-out” wear caused by high inlet velocity
common strategy for reducing wear. In some cases, it is desirable to line the bends in
order to avoid frequent replacement. In other instances, the replacement cost is
tolerated, or else the bends are fabricated of metal that is thicker, and possibly
harder, than that used in the straight lengths of pipe.
As we have seen in this chapter, wear due to slurry erosion is a complex problem,
and much remains to be learned. The opportunities for improvement in equipment
performance have not yet been exhausted, and new solutions, both theoretical and
practical, will continue to be developed for some time to come.
References
G.R. Addie, S. Bross, Modeling of slurry pump nose wear. Presented at the 15th regional phosphate
conference, Lakeland, 2000
G.R. Addie, J. Kadambi, R.V. Visintainer, Design and application, slurry pump technology.
Presented at the ASME 9th international symposium on solid-liquid flow, Houston, 2005
P. Blau, Lessons learned from the test-to-test variability of different types of wear data. Wear
376–377, 1830–1840 (2017)
S. Bross, G.R. Addie, Prediction of impeller nose wear behavior in centrifugal slurry pumps.
Presented at the 4th international conference on multiphase flow, New Orleans, 2001
J. Finnie, Erosion of surfaces by solid particles. Wear 3, 87–103 (1960)
References 441
J.M. Furlan, Particle concentration measurements in a centrifugal slurry pump using an A-scan
ultrasound technique. Dissertation, Case Western Reserve University, Cleveland, 2011
J.M. Furlan, V. Mundla, J.R. Kadambi, N. Hoyt, R.V. Visintainer, G.R. Addie, Localized particle
concentration measurement in slurry flows using a-scan ultrasound technique. Presented at the
fluids engineering division summer meeting FEDSM, Vail, 2009
J.M. Furlan, V. Mundla, J.R. Kadambi, N. Hoyt, R. Visintainer, G.R. Addie, Development of
A-scan ultrasound technique for measuring local particle concentration in slurry flows. Powder
Technol. 215, 174–184 (2012)
J.M. Furlan, M.A. Garman, J.R. Kadambi, R.V. Visintainer, K.V. Pagalthivarthi, Ultrasonic
measurements of local particle velocity and concentration within the casing of a centrifugal
pump. Presented at the ASME-JSME-KSME joint fluids engineering conference AJKFLUIDS,
Seoul, 2015
M.A. Garman, Local particle velocity measurements in slurry flow in pipes and centrifugal pumps
using ultrasound technique. Dissertation, Case Western Reserve University, Cleveland, 2015
P. Hergt, K.V. Pagalthivarthi, S. Brodersen, R.J. Visintainer, A study of the outlet velocity
characteristics of slurry pump impellers. Presented at the ASME 5th international symposium
on solid-liquid flow, Lake Tahoe, 1994
I.M. Hutchings, Tribology: Friction and Wear of Engineering Materials (Edward Arnold, London,
1992)
J.R. Kadambi, P. Charoenngam, A. Subramanian, M.P. Wernet, J.M. Sankovic, G.R. Addie,
R. Courtwright, Investigations of particle velocities in a slurry pump using PIV: Part 1, The
tongue and adjacent channel flow. ASME, J. Energy Resour. Technol. 126, 271–278 (2004)
J. Laguna-Camacho, R. Lewis, M. Vite-Torres, J. Mendez, A study of cavitation erosion on
engineering materials. Wear 301, 467–476 (2013)
J. Li, Prediction of particle rotation in a centrifugal accelerator erosion tester and the effect on
erosion rate. Wear 258, 407–502 (2005)
J.E. Miller, The SAR number for slurry abrasion resistance, in Slurry Erosion – Uses, Applications
and Test Methods, ed. by J.E. Miller, F. Schmidt, vol. 946, (ASTM STP, 1987), pp. 156–166
K.V. Pagalthivarthi, G.R. Addie, Prediction of dredge pump shell wear. Presented at the 12th World
Dredging Congress – WODCON XII, Orlando, 1989
K.V. Pagalthivarthi, F.W. Helmly, Applications of materials wear testing to solids transport via
centrifugal slurry pumps, in Wear Testing of Advanced Materials, ASTM STP 1167, ed. by
R. Divakar, (ASTM, 1990), pp. 114–126
K.V. Pagalthivarthi, R.J. Visintainer, Solid-liquid flow-induced erosion prediction in three-
dimensional pump casing, FEDSM2019-78274. Presented at the fluids engineering division
summer meeting, Vail, 2009
K.V. Pagalthivarthi, R.J. Visintainer, Finite element prediction of multi-size particulate flow
through three-dimensional channel: Code validation. J. Comp. Multiphase Flow 5(1), 57–72
(2013a)
K.V. Pagalthivarthi, R.J. Visintainer, Finite element prediction of multi-size particulate flow
through two-dimensional pump casing. Int. J. Mech. & Ind. Tech. 7(4), 88–101 (2013b)
K.V. Pagalthivarthi, P.V. Desai, G.R. Addie, Particulate motion and concentration fields in cen-
trifugal slurry pumps. Part. Sci. Technol. 8, 77–96 (1990)
K.V. Pagalthivarthi, P.V. Desai, G.R. Addie, Quasi-3D computation of turbulent flow in centrifugal
pump casings. Presented at the ASME winter annual meeting, 1991
K.V. Pagalthivarthi, J.M. Furlan, R.J. Visintainer, Wear rate prediction in multi-size particulate
flow through impeller, FEDSM2013-16192. Presented at the fluids engineering division sum-
mer meeting, Incline Village, 2013a
K.V. Pagalthivarthi, J.M. Furlan, R.J. Visintainer, Comparison of 2D and 3D predictions of erosion
wear in centrifugal slurry pump casings, FEDSM2013-16215. Presented at the fluids engineer-
ing division summer meeting, Incline Village, 2013b
442 13 Erosive Wear
K.V. Pagalthivarthi, J.M. Furlan, R.J. Visintainer, Finite element prediction of multi-size particulate
flow through three-dimensional pump casing, AJKFluids2015-31129. Presented at the ASME-
JSME-KSME joint fluids engineering conference, Seoul, 2015
K.V. Pagalthivarthi, J.M. Furlan, R.J. Visintainer, Consistent evaluation of wear coefficient from
the experiments for use in CFD Simulations, FEDSM2017-69243. Presented at the fluids
engineering division summer meeting, Waikoloa, 2017a
K.V. Pagalthivarthi, J.M. Furlan, R.J. Visintainer, Effective particle size representation for erosion
wear in centrifugal pump casings, FEDSM2017-69240. Presented at the fluids engineering
division summer meeting, Waikoloa, 2017b
S.V. Patankar, Numerical Heat Transfer and Fluid Flow (Hemisphere Publishing, London, 1980)
M.C. Roco, G.R. Addie, Analytical model and experimental studies on slurry flow and erosion in
pump casings. Presented at the 8th international conference on slurry technology, San
Franscisco, 1983
M.C. Roco, E. Reinhardt, Calculation of solid particle concentration in centrifugal pump impellers
using finite-element technique. Presented at the 7th international conference on the hydraulic
transport of solids in pipes, Sendai, 1980
M.C. Roco, C.A. Shook, Modeling slurry flow: The effect of particle size. Can. J. Chem. Eng. 61,
494–503 (1983)
M.C. Roco, P. Nair, G.R. Addie, J. Dennis, Erosion of concentrated slurries in turbulent flow, in
Liquid-Solid Flows and Erosion Wear in Industrial Equipment, ed. by M.C. Roco, vol.
13, (ASME, FED, 1984a), pp. 69–77
M.C. Roco, G.R. Addie, J. Dennis, P. Nair, Modeling erosion wear in centrifugal slurry pumps.
Presented at the 9th international conference on the hydraulic transport of solids in pipes, Rome,
1984b
M.C. Roco, P. Nair, G.R. Addie, Test approach for dense slurry erosion, in Slurry Erosion – Uses,
Applications and Test Methods, ed. by J.E. Miller, F. Schmidt, vol. 946, (ASTM STP, 1987),
pp. 185–210
A. Roudnev, R. Kosmicki, Effect of CFD modeling configuration on centrifugal slurry pump casing
wear prediction. Presented at the 17th international conference on the hydraulic transport of
solids, Cape Town, 2009
O. Schenk, K. Gartner, Solving unsymmetric sparse systems of linear equation with
PARDISO. J. Futur. Gener. Comput. Syst. 20(3), 475–487 (2004)
A. Sellgren, G.R. Addie, R.J. Visintainer, K.V. Pagalthivarthi, Prediction of slurry component wear
and cost. Presented at the 25th Ann. WEDA conference – 37th Ann. TAMU conference, New
Orleans, 2005
P.R. Spalart, S.R. Allmaras, A one-equation turbulence model for aerodynamic flows. La
Recherche Aerospatiale 1, 5–21 (1994)
G. Sundararajan, Comprehensive model for the solid particle erosion of ductile materials. Wear 149,
111–127 (1991)
H. Tian, G.R. Addie, Super corrosion-abrasion resistant white iron alloys and their applications.
Presented at the AlChE spring conference, New Orleans, 1996
H. Tian, G.R. Addie, Corrosion-wear behaviors of some high alloyed white irons and stainless
steels. AFS Trans. 97–126, 595–602 (1997)
H. Tian, G.R. Addie, Experimental study on erosive wear of some metallic materials using Coriolis
wear testing approach. Wear 258, 458–469 (2003) Presented at the international conference on
abrasive and erosive wear, Cambridge, UK
H Tian, G.R. Addie, H. Pagalthivarthi, Determination of wear coefficients for erosive wear
prediction through Coriolis wear testing. Presented at the 15th international conference on
wear of materials, San Diego, 2005
H. Tian, G.R. Addie, R.J. Visintainer, Erosion–corrosion performance of high-Cr cast iron alloys in
flowing liquid–solid slurries. Wear 267, 2039–2047 (2009)
References 443
J.J. Tuzson, J. Lee, K.A. Scheibe-Powell, Slurry erosion tests with centrifugal erosion tester, in
Liquid-Solid Flows and Erosion Wear in Industrial Equipment, ed. by M.C. Roco, vol.
13, (ASME, FED, 1984), pp. 84–87
R.J. Visintainer, D. Wolfe, The impact wear behavior of large rocks on slurry pump materials and
equipment. Presented at the Western Dredging Association Dredging Summit & Expo,
Houston, 2015
R.J. Visintainer, J.M. Furlan, G. McCall, P. Springer, CFD simulation and scale model testing of a
new dredge pump design. Presented at the 22nd World Dredging Congress – WODCON XXII,
Shanghai, 2019
R.J. Visintainer, M. Garman, K.V. Pagalthivarthi, Parametric study of slurry pump wear and
operating cost. Presented at the 23rd World Dredging Congress – WODCON XXIII,
Copenhagen, 2022
F.M. White, Viscous Fluid Flow, 2nd edn. (McGraw-Hill, New York, 1991)
W. Wiedenroth, An experimental study of wear of centrifugal pumps and pipeline components, in
Liquid-Solid Flows and Erosion Wear in Industrial Equipment, ed. by M.C. Roco, vol.
13, (ASME, FED, 1984a), pp. 78–83
W. Wiedenroth, Wear tests executed with a 125 mm I.D. loop and a model dredge pump. Presented
at the 9th international conference on the hydraulic transport of solids in pipes, Rome, 1984b
W. Wiedenroth, Wear of solids-handling centrifugal pump impellers. Presented at the 11th inter-
national conference on the hydraulic transport of solids in pipes, Stratford-upon-Avon,
UK, 1988
Y. Xie, J. Jiang, K. Tufa, S. Yick, Wear resistant of materials used for slurry transport. Wear
332–333, 1104–1110 (2015)
Chapter 14
Pump Selection and Cost of Ownership
Centrifugal slurry pump selection begins according to the same approach commonly
used for clear fluid pumps: A range of pumps will be available, each having its own
head-flow characteristic. For each pump, the rotational speed is determined at which
its head characteristic intersects with the head and flow required by the application.
This is done via the scaling laws developed in Sect. 8.1.
Next, the hydraulic and mechanical limitations of each pump (as defined by the
pump manufacturer) are checked against the operating conditions, taking into
account the required operating speed and the density of the mixture being pumped.
These limitations may include the following:
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 445
R. Visintainer et al., Slurry Transport Using Centrifugal Pumps,
https://doi.org/10.1007/978-3-031-25440-6_14
446 14 Pump Selection and Cost of Ownership
Hydraulic Limitations
• Pump NPSHR compared to system NPSHA, as covered in Chaps. 8 and 11.
• The ratio of the application flow rate to the pump’s best efficiency flow rate (Q/
QBEP). Many pump types are limited to operate within an allowable range of this
ratio. Specific recommendations for slurry pumps are addressed in Sect. 14.2.2.
• Additional pump-specific limitations on head or flow rate may also be imposed.
Mechanical Limitations
• Pressure (maximum operating and required test pressures)
• Operating temperature
• Chemical resistance (against pH, chlorides, etc.)
• Shaft fatigue stress
• Shaft deflection at the sealing locations or other critical points
• Shaft resonant speeds (lateral and torsional)
• Bearing life
• Impeller, bearing, and seal limiting speeds
• Available driver power and speed (if pre-existing)
• Any other pump-specific limitations
Pumps exceeding any of the defined limitations are eliminated, and the remaining
selections are often sorted according to their size, operating efficiency, or Q/QBEP
ratio. Next, the power, maintenance, and capital costs of these pumps are evaluated
against the technical, logistical, and financial priorities of the operator, and a final
selection is made. As operators’ priorities and needs will vary, there will be no
universal best selection for any given application.
In a slurry application, neither the most efficient nor the lowest cost design will
automatically be chosen, due to the important role played by wear performance and
other slurry-specific considerations.
Slurry pump selection will differ from clear fluid pump selection primarily due to
extra limitations imposed by the effects of the slurry. The most common of these are:
• The maximum solids size that the pump can pass without clogging
• The effects of the slurry viscosity and solids on pump performance (head,
efficiency, and NPSHR)
• The effects of erosive wear on pump component service life
The maximum solids size is fairly straightforward, although some margin should
be allowed between the size of the largest solids and that of the pump passage size.
This margin will vary with experience and type of solids. Typical margins call for a
pump passing size that is 1.5–2 times the largest solid to be encountered.
14.1 Basic Principles of Centrifugal Slurry Pump Selection 447
The effect of solids on pump performance will typically require the pump to run at
a higher speed, requiring more power and NPSHA. The procedures for determining
and correcting for the pump solids effect are covered in Chap. 9.
The topic of erosive wear is more complex and is treated at length in Chap. 13. As
may be imagined, it is of considerable importance in slurry pump selection. Whereas
mechanical considerations (such as rotational speed) and NPSHR tend to limit clear
fluid pump selection, wear considerations often impose more stringent limitations on
slurry pump selection. These limitations push the selection towards pumps that are
larger, slower running, and composed of more wear-resistant materials, all of which
tend to increase the cost of the pump. Therefore, finding the best balance between
pump wear performance and cost is essential to optimizing the selection of a slurry
pump. Based on the treatment in Chap. 13, together with much practical industrial
experience, a number of guidelines have been developed for addressing erosive wear
during the slurry pump selection process. These are discussed in detail in Sect. 14.2.
If the system head is more than one pump can handle, then multiple pumps in series
may be used. This setup allows the head to be divided among the pumps. For a clear
fluid application, the number of pumps would normally be minimized by taking each
pump as close as possible to its maximum allowable head, thus reducing the total
capital cost. Furthermore, a multi-stage pump may be used (i.e. having multiple
stages of impellers & diffusers within a single machine). With slurry pumps, multi-
stage machines are rarely used, due to erosive wear considerations. Furthermore, the
question of dividing head among pumps in a multi-pump system becomes more
complex, since the head per pump will affect each pump’s wear life. In severe-duty
applications with no installed spares, the replacement cycle for the wear parts may
determine the maintenance cycle of the entire process, incurring what is known as a
downtime cost attributable to the pump. In other words, using more pumps with less
head per stage might cost more in capital, but can sometimes provide a quick
payback by reducing future production losses due to outages. This concept is
discussed in greater detail in Sect. 14.3.
In a similar way, multiple pumps in parallel may be used if the flow rate exceeds
the available pump capacity, although this setup is less common in slurry applica-
tions. See Chap. 11 for further discussion of the operational issues associated with
multiple pump systems.
448 14 Pump Selection and Cost of Ownership
In addressing erosive wear considerations, the first step is to define the erosive
potential of the slurry itself. This is the purpose of the Slurry Service Class chart
given in Fig. 14.1. Based on the median particle size (d50) and slurry specific gravity
(Sm), a slurry service class is determined, ranging from light duty (Class 1) to severe
duty (Class 4).
This chart assumes a “normal” particle size distribution, defined as following a
Weibull distribution curve and having a d85 particle passing size equal to 2.3 times
the d50 size. For other particle size distributions, the d50 size used when reading the
chart is corrected as follows:
ðd50 þ d85 Þ
d50,corrected = ð14:1Þ
3:3
The chart also assumes a particle abrasivity equivalent to that of typical silica-
based sand. For particles with other abrasivity, both the Sm and d50 values used when
reading the chart are corrected as follows:
Fig. 14.1 Slurry service class chart for Ss = 2.65 silica-based solids
14.2 Wear Considerations 449
1.6
1.4
1.2
1.0
Abrasivity (Apart)
0.8
0.6
0.4
silica sand
0.2
0.0
0
50
100
150
200
250
300
Miller Number (per ASTM G75)
Fig. 14.2 Particle abrasivity as a function of Miller number for pump wear applications
Sm,corrected = Sm 0:4Apart þ 0:6 ð14:2Þ
where Apart is the particle abrasivity for erosive wear relative to typical silica-based
sand. The typical range of the relative abrasivity encountered in mining and dredging
applications runs from 0.6 for limestone to 1.5 for some of the harder metallic ores.
More detailed guidance about particle abrasivity may be found in the Miller Number
Slurry Abrasivity Standard ASTM G75-15 (2021) and Fig. 14.2.
Once the service class has been determined, recommended limits are applied to the
pump operating conditions based on the service class number. The two most
important limits are for pump head (H ) and the ratio of operating flow rate to the
best efficiency flow rate (Q/QBEP). These are the two most significant variables of
pump operation affecting wear, as was seen in the parametric wear study of Sect.
13.9.
The recommended limits on head for metal-lined slurry pumps are given in
Table 14.1. Limits are given for both head and impeller peripheral velocity, which
is closely related to head. Either limit (or both) may be used, as the delineation is
not precise. For pumps lined with natural rubber, constant limits of 40 m head and
450 14 Pump Selection and Cost of Ownership
Table 14.1 Recommended pump head and impeller peripheral velocity limits for metal-lined
slurry pumps as a function of slurry service class
Slurry service class
Recommended limits for acceptable wear 1 2 3 4
Pump head (m) 105 73 55 40
Impeller peripheral velocity (m/s) 46 38 33 28
Table 14.2 Recommended pump relative operating flow rate (Q/QBEP) as a function of slurry
service class
28 m/s impeller peripheral speed are recommended for all service classes, since the
primary limitation is heat build-up due to rubber vibration rather than to wear. (See
“Elastomers” in Sect. 13.3) Other, harder elastomers, such as neoprene, may tolerate
limits that are 10–20% higher than the natural rubber value. Polyurethanes can
achieve limits similar to those for a Class 2 or Class 3 service, depending on
hardness.
Table 14.2 gives the recommended limits on operating flow rate. These limits
vary according to the shape of the slurry pump casing, with ranges shifting to higher
flow rates as the casing type shifts from annular to near volute. In the previous
chapter, Fig. 13.23 shows pictorial description of the casing types, as well as a
graphical representation of Table 14.2. Note that the ranges of recommended flow
rate become narrower as the slurry class increases and the head approaches the
recommended maximum from Table 14.1.
Figure 14.3 combines the limits of Tables 14.1 and 14.2 for the semi-volute
casing design. The reduction in the allowable flow rate range with increased head is
due to the increased risk of hydraulic and mechanical vibrations, as well as increased
wear at higher heads. Additional limits may be applied to specialized slurry pump
designs, in particular:
14.2 Wear Considerations 451
120
for
semi-volute
100 casing type
80 Slurry Class:
Class 1
H (m)
60
Class 2
40 Class 3
Class 4
20
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4
Q/QBEP
Fig. 14.3 Combined plot of pump head and Q/QBEP limit recommendations for the semi-volute
casing type
• For pumps with 3-vane impellers: Replace Class 1 limits with Class 2 limits, due
to the potential for excessive vibration at higher heads.
• For froth pumps: Replace Class 1 limits with Class 2 limits, due to the potential
for air-lock and cavitation at higher heads.
• For pumps with specific speed Ns > 60: Use the Q/QBEP limit for the 100% head
condition for all heads, due to increased potential for unsteady hydraulic flow.
• For single-volute vertical pumps: Use the Q/QBEP limit for the 100% head
condition for all heads, due to the increased potential for excessive vibration.
While these operational limits are in agreement with the theoretical treatment of
wear given in Chap. 13, their exact values are largely based on practical experience.
Their status as guidelines, rather than fixed rules, should be recognized. That being
said, the service class chart lines can be understood as representing contours of
approximately constant wear potential. Maintaining operations within the
recommended limits, based on the service class, will help ensure that acceptable
pump wear life is achieved. The term “acceptable” in this context is subjective; in
actual practice, the following outcomes are typical for pumps operating near the
maximum head for each class:
• 24–48 months wear life for a Class 1 application
• 12–24 months wear life for a Class 2 application
• 6–12 months wear life for a Class 3 application
• 3–6 months wear life for a Class 4 application
452 14 Pump Selection and Cost of Ownership
Note that wear life decreases with increasing service class, even though the
operating limits become more stringent. This decrease highlights the strong influ-
ence that the slurry properties of particle size and concentration have on pump wear.
In the end, economic considerations will determine the acceptable wear life, and
thereby, the parameters that should govern pump selection. The next section inves-
tigates these concepts.
As stated previously, the limits recommended in this section are based on both
theory and experience obtained by engineers at GIW Industries over the past
40 years. These limits have also been adopted into the Hydraulic Institute Standard
for Centrifugal Slurry Pumps (ANSI/HI 12.1-12.6 2021), which contains other
valuable guidelines for design and application. It is recommended reading for any
engineer working with centrifugal slurry pumps.
In the following analysis, we will again assume that the pipeline system head
characteristic, slurry composition, and desired flow rate have already been deter-
mined and will focus on determining the costs of pump operation. We will, however,
also consider in our analysis how changes in system design conditions, particularly
in the slurry properties of particle diameter (d ), density (ρs), and volumetric concen-
tration (Cv), may affect pump operating cost, since their effects on the pump may
vary from those they have on the pipeline system.
will be expressed as cost per ton of solids delivered, so they are linked to the value of
the solids transported.
Power Cost
The power cost is determined by the hydraulic power generated, the pump effi-
ciency, and the cost of power, as follows:
where $P = the power cost in ($/kWh); Sm, Cvd, and ρs are the slurry mixture specific
gravity, volumetric concentration of solids, and solids density (kg/m3); Q, H, and η
are the pump flow rate (m3/s), head (m slurry), and efficiency, including the derate
effect of the solids; and TPH is the delivered tons of solids per hour. In the following
analysis, we will assume a power cost of 0.10 $/kWh.
Wear Parts Replacement Cost
The cost of replacement wear parts depends on the part purchase price, the calculated
wear rate, and the controlling wear thickness, as follows:
_ WP =ðTPH 1000T WP Þ
Wear Parts Cost =tonneÞ = WP W ð14:6Þ
where $WP = the wear part purchase price ($), W _ WP = the controlling wear rate
(μm/h), and TWP is the controlling wear thickness. This wear part cost must be
considered for each major part in turn, and it typically consists of a casing, impeller,
and side liner(s). This cost is more difficult to predict than the power cost, since the
precise factors of wear rate and the controlling wear thickness are not often known
with precision. In cases where some experience in a similar application exists, the
expected wear rate can be estimated by calibrating the equations given in Sect. 13.9
to the known condition, and then modifying the parameters to reflect the new
condition and calculating the new wear rate. Of course, the closer the known
condition is to the new condition, the more accurate the estimate will be. When in
doubt, it is best to be somewhat conservative with these estimates, allowing some
buffer for uncertainty by applying a safety factor to the wear rate.
Determination of the controlling wear thickness is also uncertain. Parts may not
wear evenly (and are not always allowed to fully “wear out” before being replaced),
and maintenance cycles must be aligned across many pieces of equipment for
efficiency of operations and reduction of downtime. A reasonable rule of thumb is
to use 50% of the full thickness for impellers and liners and 75% of the full thickness
for casings. The smaller percentage is recommended for the impeller and liner, as
these more often fail due to localized wear patterns that develop over time as the part
wears out. Impellers and liners are also more likely to be retired early due to fixed
maintenance cycles.
454 14 Pump Selection and Cost of Ownership
In the following analysis, the wear rate equations of Sect. 13.9 were used outright.
For part thicknesses, correlations were developed for each part type based on the
actual thicknesses of the pumps analyzed, in order to smooth out the effects of
unusually thick or thin parts on the average result. While this approach represents an
obvious simplification, the goal is to examine trends and incremental effects. A
similar approach was taken with regard to parts pricing, and the same comment
applies regarding the simplification this represents. The total wear cost was calcu-
lated by summing up the individual part costs for suction liner, casing, impeller, and
drive-side liner. The drive-side liner was assumed to have the same wear life as the
casing and the same price as the suction-side liner.
Maintenance and Capital Costs
For simplicity in the present study, maintenance costs were assumed to be equal to
10% of the total wear parts’ cost per ton, and capital costs were taken as 10% of the
purchase price of the complete pump per 6000 operating hours.
For an accurate analysis of these costs in the modeling of a specific installation,
more detailed and realistic cost models can, and should, be used. In particular,
maintenance costs will vary greatly from site to site and may include mobilization
fees. A scarcity of qualified workers may also factor into the cost of maintenance.
With regard to capital costs, we have considered here only the bare shaft pump, but a
specific analysis might also include the cost of the drives, along with other facilities
required to support the pump.
Combining the above costs, a total operating cost per ton of solids delivered was
calculated for the same collection of pump designs and operating conditions used in
Sect. 13.9. The strongest correlating variables for predicted cost were the pump-
specific speed (Ns), expressed in customary metric units (See Sect. 8.2.3), and the
pump size, as represented by the impeller diameter (D), expressed in meters.
Figures 14.4 and 14.5 show results for the parametric study’s baseline operating
condition of H = 50 m, ρs = 2650 kg/m3, Cv = 0.2, and deff = 500 μm, where deff is
the effective particle size for wear, taken as the average of the d50 and d85 particle
sizes.
A number of observations may be made from these figures. Wear costs vary
considerably and increase sharply as the parameter (D Ns) decreases. This effect is
not surprising, as the wear models reveal a size effect for most of the wear parts, and
the economies of scale associated with running larger equipment are well
established. The increase in cost that accompanies a reduction in specific speed,
however, runs counter to the experience that lower specific speed pumps often wear
at lower rates. But while decreased wear rate may reduce downtime and increase the
time between maintenance (cost effects that are not included in the current analysis),
lower specific speed machines also tend to be larger and more expensive. This
14.3 Economic Considerations 455
0.100
Operating Cost ($/tonne)
0.010
0.001
0.000
0 20 40 60 80 100 120 140
D (m) · NS
Fig. 14.4 Operating cost of wear parts for the baseline operating condition. (Visintainer et al. 2022)
0.100
Operating Cost ($/tonne)
0.010
0.001
0.000
0 20 40 60 80 100 120 140
D (m) · NS
Fig. 14.5 Total pump operating cost for the baseline operating condition. (Visintainer et al. 2022)
456 14 Pump Selection and Cost of Ownership
counterbalance appears in the analysis, especially with smaller pumps. The effect
“flattens” as the (D Ns) parameter increases, beginning at a value near 30, which
represents (for example) a specific speed of Ns = 30 for a 1 m impeller, or Ns = 20
for a 1.5 m impeller. The curves for suction liner and impeller cost are flatter than for
the casing, since they are both more sensitive to increases in wear with specific
speed. This effect cancels out some of the gains from an increasing pump size.
It should be emphasized here that the (D Ns) parameter will not vary greatly
between pumps selected for any given application, since as one moves to a higher
specific speed design for any given flow and head, the impeller diameter will
decrease in order to maintain the same ratio of Q/QBEP. As a result, these two effects
will roughly cancel each other out. However, as seen in Fig. 13.19, impeller and liner
wear rate will increase as specific speed increases, and the mean time between part
replacement for a higher specific speed pump will therefore be shorter. This shorter
time may not have a strong effect on operating cost as calculated here, since the parts
are also smaller; however, if there are additional costs incurred by a short mainte-
nance cycle that are not included in the present analysis (for example, downtime
costs; see Sect. 14.3), the difference can become important.
Compared to the other costs, power cost is relatively flat, as it is affected only by
the pump efficiency. Efficiency may, of course, decrease as the pump wears, but we
do not consider that here. Power dominates the cost for larger pumps, while wear
parts cost reaches a similar magnitude for smaller pumps. This balance will shift as
the severity of the abrasive service changes. For example, in Fig. 14.6, we see the
results for a more severe duty of deff = 1000 μm and Cv = 0.4. The wear parts and
maintenance costs have shifted upwards considerably relative to Fig. 14.5, while the
power cost has changed only marginally, increasing at smaller pump sizes due to a
reduced pump efficiency, but actually decreasing with larger pumps due to the
beneficial effect of transporting less water at a higher solids concentration.
In Fig. 14.7, the effects of the various operating parameters are examined, each in
turn, while the remaining parameters are held at their baseline values. A number of
interesting relationships can be seen in this plot.
• Both power and wear costs increase with head, since both energy consumption
and the associated wear-causing energy dissipation increase together.
• In the case of particle size, there is very little effect on the power consumption,
limited to the solids effect on pump performance. However, the effect on parts
wear is considerable.
• In the case of solids concentration, this arrangement is reversed. Energy effi-
ciency increases greatly as concentration increases because less water is being
transported. Wear is also increased by concentration, but so is the throughput,
such that the wear costs (calculated according to $/ton) balance out.
• The Q/QBEP ratio affects the power cost at low values due to the degradation in
pump efficiency. Otherwise, the effect is mild. Wear costs are affected more
strongly, with a minimum near Q/QBEP = 0.75, and maximums at the lowest and
highest ratios.
14.3 Economic Considerations 457
0.100
Operating Cost ($/tonne)
0.010
0.001
0.000
0 20 40 60 80 100 120 140
D (m) · NS
Fig. 14.6 Total pump operating cost for a modified baseline condition having a more erosive
slurry. (Visintainer et al. 2022)
Considering all of the results for variable operating conditions and other geomet-
ric ratios, generalized correlations for total cost of ownership (TCO) in US dollars
($) per ton were developed. Because the power costs are relatively flat compared to
the costs of wear and maintenance, the results are highly non-linear and do not easily
fit into a single equation. It was found necessary to perform a piece-wise correlation
for different ranges of the fitting parameter (D Ns), as follows:
For (D Ns) > 25:
0:9 ρ - 0:45
H deff 0:3 Cv - 0:6
TCO = 0:125 S
50 500 0:2 2650
h i0:75 ð14:7Þ
Q
1 þ abs - 1:1 ðD N s Þ - 0:25
QBEP
0:8 ρ - 0:45
H deff 0:7 Cv - 0:4
TCO = 0:8 S
50 500 0:2 2650
h i1:2 ð14:8Þ
Q
1 þ abs - 0:95 ðD N s Þ - 0:85
QBEP
458 14 Pump Selection and Cost of Ownership
deff (mm) =
deff 0.2 0.4 0.8 1.5 2.5
1.000 1 1.000
Wear Parts Cost ($/tonne)
Power Cost ($/tonne)
Fig. 14.7 The effects of the various operating parameters on power, wear parts, and total operating
cost. Each parameter is varied singly relative to the baseline operating case. (Visintainer et al. 2022)
0:65 ρ - 0:45
H deff 1:2 Cv - 0:2
TCO = 2:2 S
50 500 0:2 2650
h i2 ð14:9Þ
Q
1 þ abs - 0:9 ðD N s Þ - 1:25
QBEP
14.3 Economic Considerations 459
The differences in the ranges are seen in the exponent of the (D Ns) parameter,
which is included in each equation and increases in magnitude as the parameter value
decreases. Notable variations in dependencies of the other variables are also seen. In
particular:
• The importance of particle size and Q/QBEP ratio increases as the (D Ns)
parameter decreases. At the same time, the optimal value of the Q/QBEP ratio
shifts from 1.1 to 0.9. This effect is due to the increasing influence of the
wear cost.
• The importance of head and concentration increases as the (D Ns) parameter
increases, due to the increasing influence of the power cost.
• For the larger pumps, where (D Ns) > 25, cost increases roughly in direct ratio to
head (i.e., H0.9). This means that using more pumps at lower head will cost
roughly the same as using fewer pumps at higher head (not including any effect of
downtime). At the smaller end of the parameter range, cost varies by the 0.65
power of head, so that the lowest cost will be afforded by using the smallest
number of pumps operating at the highest allowable head each.
As mentioned previously, costs and prices will vary with place and time, altering
the constants in these equations. Furthermore, broader or narrower assumptions may
be made regarding what is included in the calculation. However, by calibrating the
wear correlations and other calculation methodologies described above, the reader
may construct their own analysis using the wear thicknesses, power costs, parts
pricing, and financial details relevant to their immediate circumstances.
In the previous economic analysis, we have assumed that no extra costs beyond
those required for pump maintenance activities are incurred whenever pump parts
wear out and must be replaced. This is a reasonable assumption in cases where
installed spare pumps are available or where the pump achieves a normal mainte-
nance cycle determined by other equipment in the system. However, in many mining
and dredging operations, it is not uncommon for there to be no installed spares, due
to the cost and size of the equipment. Furthermore, it is not uncommon for the pump
to be the highest-wearing component in the slurry system. The centrifugal pump
adds large amounts of energy to the slurry flow by temporarily increasing its velocity
within a rotating impeller, and wear is generally proportional to the third power of
velocity. The 10–20% of this energy that is dissipated within the pump will, in part,
be expended in producing wear.
In cases where the pump is the primary cause of a plant stoppage, a pump
downtime cost is incurred, which we define here as the lost revenue due to lost
production when pump downtime is the primary cause of a production stoppage.
Downtime costs can be significant, as large operations may have production revenue
values exceeding $100,000 (US) per hour. In some cases, the downtime cost
460 14 Pump Selection and Cost of Ownership
incurred during a shutdown will exceed all of the other pump costs (power, wear,
maintenance, and capital) summed over the entire run cycle leading to the shutdown
period.
Downtime costs can vary widely. Whenever they occur, they should be included
in the cost-of-ownership analysis and will often shift the analysis towards a higher
cost for pump wear and capital in order to reduce or eliminate the cost of downtime.
References
ANSI/HI 12.1-12.6, American National Standard for Rotodynamic Centrifugal Slurry Pumps
(Hydraulic Institute, Parsippany, 2021)
ASTM G75-15, Standard Test Method for Determination of Slurry Abrasivity (Miller Number) and
Slurry Abrasion Response of Materials (SAR Number) (ASTM International,
W. Conshohocken, 2021)
R. Visintainer, M. Garman, K. Pagalthivarthi, Parametric study of slurry pump wear and operating
cost. Presented at the 23rd World Dredging Congress – WODCON XXIII, Copenhagen 2022
(2002)
Appendix: VSCALC Function
The function VSCALC can be called by MATLAB scripts to calculate the settling
slurry deposition limit velocitiesVsm and Vs, as discussed in this book. For the
calculation of Vsm, the function uses the equations presented in Chaps. 2, 4, and 7,
and references to the equations are given in the script comments. For the calculation
of Vs, i.e., for the quantification of the effect of solids concentration on the deposition
limit velocity, the script uses equations presented in the 2nd Edition of the book.
The function input includes the pipe parameters of internal diameter (D), wall
roughness (ε), and inclination angle (θ); the solids properties of particle diameter (d),
density (ρs), and terminal settling velocity of a spherical particle (vts); the carrier fluid
properties of density (ρf) and viscosity (μ); the delivered concentration of solids
(Cvd), the bed concentration (Cvb), and the sliding-bed friction coefficient (μs).
© The Editor(s) (if applicable) and The Author(s), under exclusive license to 461
Springer Nature Switzerland AG 2023
R. Visintainer et al., Slurry Transport Using Centrifugal Pumps,
https://doi.org/10.1007/978-3-031-25440-6
462 Appendix: VSCALC Function
% VSCALC predicts deposition limit velocities Vsm and Vs for settling slurries in
horizontal and inclined pipes
% INPUTS
% D = pipe inner diameter (m)
% eps = pipe wall roughness (m)
% vts = terminal settling velocity of spherical particle of diameter d_mm (m/s)
% d_mm = equivalent diameter of solid particle (mm)
% ros = density of solid particle (kg/m3)
% rof = density of carrying liquid (kg/m3)
% mu = dynamic viscosity of carrying liquid (Pa.s)
% Cvd = delivered volumetric concentration of solids (-)
% Cvb = bed volumetric concentration (-)
% mus = friction coefficient of sliding bed (-)
% thetaDeg = pipe inclination angle (deg)
% CONSTANTS
g = 9.81;
row = 1000; % standard water density
% COMPUTATIONS
theta = thetaDeg*pi/180;
sintheta = sin(theta);
costheta = cos(theta);
Sf = rof/row;
Ss = ros/row;
nu = mu/rof;
d = 0.001*d_mm;
Cds = 4/3.*(ros-rof)./rof.*g.*d./vts.^2; % drag coefficient for spherical particle
% horizontal flow
FL = Vsm./sqrt(2.*g.*(Ss./Sf-1).*D);
Ref = D.*Vsm./nu;
ff = 0.25./(log10(eps./3.7./D+5.74./Ref.^0.9)).^2; % Eq. 2.43
Vsmmax_est = (0.018./ff).^0.13.*sqrt(2.*g.*(Ss./Sf-1).*D); % estimated Vsm,max
% inclined flow
DeltaD = 0.75.*theta-0.50.*(0.6366.*theta).^2./(1-0.6366.*theta);
if theta < 0
DeltaD = 0.75.*theta-0.02.*(2.29*theta).^2./(1-2.29.*theta);
end
if Vsm < 0;
Vsm = 0;
FL = 0;
end
% velocity Vs -----------------------------------------------
Cr = Cvd./Cvb;
Crm = 0.16.*D.^0.4./(d_mm.^0.84.*(mus.*(Ss./Sf-
1)./0.66).^(0.55.*0.3))./(costheta+2.145.*sintheta).^0.17;
if Crm < 0.05;
Crm = 0.05;
end;
464 Appendix: VSCALC Function
alpha = log(0.333)./log(Crm);
beta = log(0.667)./log(1-Crm);
VsVsm = Cr.^alpha.*(1-Cr.^alpha).^2;
Vs = 6.75.*VsVsm.*Vsm;
% OUTPUTS
Vsm_VSCALC = Vsm;
Vs_VSCALC = Vs;
Index
© The Editor(s) (if applicable) and The Author(s), under exclusive license to 465
Springer Nature Switzerland AG 2023
R. Visintainer et al., Slurry Transport Using Centrifugal Pumps,
https://doi.org/10.1007/978-3-031-25440-6
466 Index
Centrifugal slurry pump (cont.) Downtime, 318, 404, 420, 447, 453, 454, 456,
reverse flow, 337–338 459, 460
shutdown and start-up, 318, 322, 332, 333, pump downtime cost, 459
339, 349 Drag coefficient, 21, 45, 278
shut head, 337, 339 Dredging, 5, 59, 154, 209, 227, 251, 253, 301,
single-wall design, 256 322, 345, 346, 355, 411, 459
slurry pump operating limits, 445, 449 ladder pumps, 322, 323, 346, 347, 356
slurry pump selection, 445–447 Drive train, 336–338, 357, 389
solids effect, 257, 259, 265, 272, 304, 306, See also Centrifugal slurry pump
307, 310, 447, 456 Durand equation, 116
specific speed, 239–249, 252, 261, 308, Dynamic yield stress, 376
322, 323, 330, 426, 431–435, 451, 454,
456
speed matching, 318 E
suction conditions, 319, 323 Elastomers, 256, 405, 406, 417, 450
suction performance, 348 Electrical resistance tomography (ERT), 183,
suction piping, 249, 331, 340–348, 350, 394
351, 354 Energy grade line, 29
suction specific speed, 252 Equivalent liquid, 115, 278, 294, 302–303, 311
surface roughness, 237 Equivalent liquid model (ELM), 117–119,
theoretical head, 241–243 226, 266
torque speed curve, 330 Equivalent liquid point, 305, 311
variable speed, 318, 319, 324, 332, 355, 357 Equivalent sand grain roughness, 37
viscous losses, 242 Erosive wear, 252, 401, 416, 417, 446–449
Ceramics, 256, 408, 410 See also Wear
Characteristic velocity, 72, 148, 381
Choke station, 321
Complex slurries, 67, 93, 310, 312 F
Composite materials, 408, 409 Flow measurements, 370
Computational fluid dynamics (CFD), 93, 141, See also Instrumentation
244, 254, 419, 420, 423, 425, 427, 428 Flow regime, 67
Concentration profile, 66, 83, 84, 181, 184, Flow structure, 89
188, 213 4-component model (4CM), 130, 216, 273, 277,
Conservation equation, 85 279
Constitutive equation, 11, 14 component volume fraction, 280
Contact friction, 71, 76 pump solids effect, 278–282
Continuity equation, 24 Friction factor, 33, 36, 40, 91, 309
Coriolis flow meter, 378–379 friction factor interfacial, 89
Corrosion-erosion, 411 Frothy mixtures, 16, 349–354, 451
See also Wear froth factor, 353, 354
vented pump design, 353
Fully developed flow, 24
D Fully rough, 34, 37
Delivered concentration, 60, 61, 102, 105, 211 Fully stratified flow, 66–68, 70, 84, 115, 116,
Density, 311, 370, 387 120, 121, 123, 124, 131, 135, 279, 280,
See also Solids density 284–285, 308, 334
Density measurement, 217
Density waves, 101, 140, 141, 334
Deposition limit velocity, 72, 96, 99–101, 103, G
105, 108, 123, 135, 218–220, 304, 307, GIW Hydraulic Lab, xi, xv
332, 334, 461 GIW Slurry Course, xi, xv
Deposition limit velocity maximum, 101 Graham and Pullum method, 291–294
Downhill flow, 321
Index 467
P
M Partially stratified flow, 66, 84, 124, 223
Magnetic flow meters, 377, 378, 381 Particle attrition, 144
Magnetic resonance imaging (MRI), 395 Particle diameter, 19, 131
Mass median diameter, 19 Particle image velocimetry (PIV), 392
Miller number, 449 Particle Reynolds number, 44
See also Wear testing Particle settling, 67
Mineral processing, 2, 108, 253, 346, 353 Particle shape factor, 21, 23, 48, 49
468 Index
T
Terminal velocity, 43 W
Thickened tailings, 53, 93, 268, 274, 287–288 Walker and Goulas method, 290–291
Thickener discharge, 186 Water hammer, 249, 256, 322, 331, 333,
Thixotropy, 13, 16 335–336, 338–340, 346, 382
Total cost of ownership (TCO), 318, 454, 457 cavitation, 335
Transient conditions, 212, 318, 322, 329–333 Wear, 399
Transitional rough, 37 cavitation wear, 411
Transition velocity, 310 corrosion-erosion, 410, 411
Turbulent core, 32 impact wear, 401, 408, 415, 418, 424
Turbulent flow, 30, 31, 34, 37, 67, 69, 93, sliding-abrasion, 401, 403, 404, 407, 413,
310, 374 414, 417, 425
non-Newtonian, 168–172, 186–190 slurry abrasivity, 400, 412, 413, 427, 448,
Turbulent suspension, 68, 69, 82, 94 449
Two-layer model, 83, 96, 120, 224 slurry pump wear, 426–434, 437
wear mechanisms, 404, 413
Wear coefficients, 412, 413, 415–419, 423,
U 425, 427
Unified layered model, 93, 139, 225 Wear modeling, 400, 416, 417
Unsteady flow, 140–141, 392 See also Numerical modeling
Wear-resistant materials, 404–410
ceramics, 408 (see also Ceramics)
V composites, 408–410 (see also Composite
Vapor pressure, 249 materials)
Velocity, 171 elastomers, 405–408 (see also Elastomers)
celerity, 335 polyurethane, 407–408
friction velocity, 31 (see also Polyurethane)
hindered settling velocity, 212 rubber, 406–407 (see also Rubber)
470 Index
Wear-resistant materials (cont.) White iron, 256, 404, 408, 410, 413, 417, 424,
white iron, 404–405 (see also White iron 425, 427
Wear testing, 411, 412 Wilson & Thomas method, 94, 170
Coriolis wear test, 414 Wilson–Tse chart, 220
impact wear test, 415, 424 Worster–Denny formula, 222
Miller machine, 413