12.02.2013 Views

slurry systems handbook baha e. abulnaga, pe

slurry systems handbook baha e. abulnaga, pe

slurry systems handbook baha e. abulnaga, pe

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

SLURRY<br />

SYSTEMS<br />

HANDBOOK<br />

BAHA E. ABULNAGA, P.E.<br />

Mazdak International, Inc.<br />

McGRAW-HILL<br />

New York Chicago San Francisco Lisbon London Madrid<br />

Mexico City Milan New Delhi San Juan Seoul<br />

Singapore Sydney Toronto


Cataloging-in-Publication Data is on file with the Library of Congress.<br />

Copyright © 2002 by The McGraw-Hill Companies, Inc. All rights reserved. Printed in the United<br />

States of America. Except as <strong>pe</strong>rmitted under the United States Copyright Act of 1976, no part of this<br />

publication may be reproduced or distributed in any form or by any means, or stored in a data base or<br />

retrieval system, without the prior written <strong>pe</strong>rmission of the publisher.<br />

123457890 DOC/DOC 0765432<br />

ISBN 0-07-137508-2<br />

The sponsoring editor for this book was Larry S. Hager and the production<br />

su<strong>pe</strong>rvisor was Sherri Souffrance. It was set in Times Roman by Am<strong>pe</strong>rsand<br />

Graphics, Ltd.<br />

Printed and bound by R. R. Donnelley and Sons, Co.<br />

This book was printed on recycled, acid-free pa<strong>pe</strong>r<br />

containing a minimum of 50% recycled de-inked fiber.<br />

McGraw-Hill books are available at s<strong>pe</strong>cial quantity discounts to use as premiums and sales promotions,<br />

or for use in corporate training programs. For more information, please write to the Director of<br />

S<strong>pe</strong>cial Sales, McGraw-Hill Professional Publishing, Two Penn Plaza, New York, NY 10121-2298.<br />

Or contact your local bookstore.<br />

Information contained in this work has been obtained by The McGraw-Hill Companies, Inc.<br />

(“McGraw-Hill”) from sources believed to be reliable. However, neither McGraw-Hill nor its authors<br />

guarantee the accuracy or completeness of any information published herein, and neither<br />

McGraw-Hill nor its authors shall be responsible for any errors, omissions, or damages arising<br />

out of use of this information. This work is published with the understanding that McGraw-Hill<br />

and its authors are supplying information but are not attempting to render engineering or other<br />

professional services. If such services are required, the assistance of an appropriate professional<br />

should be sought.


In memory of my father, Dr. Sayed Abul Naga,<br />

and in dedication to my mother, Dr. Hiam Aboul Hussein,<br />

who devoted their lives to comparative literature<br />

as authors and translators. May their efforts contribute<br />

to a better understanding among mankind.<br />

And to my children Sayed and Alexander<br />

for filling my life with joy and happiness.


BAHA ABULNAGA, P.E., obtained his Bachelor of Aeronautical Engineering in<br />

1980 from the University of London and his Masters in Materials Engineering in 1986<br />

from the American University of Cairo, Egypt. The first years of his professional career<br />

were devoted to the adaptation of air cushion platforms to desert environments, as well as<br />

the development of renewable energy <strong>systems</strong>. In 1988, he joined CSIRO (Australia) as a<br />

scientist. There he conducted research on complex multiphase flow for the design of<br />

smelting furnaces.<br />

Since 1990, he has been active in design of rotating equipment, pumps, and <strong>slurry</strong><br />

pi<strong>pe</strong>lines and processing plants. His career has been a balanced mixture of design of<br />

equipment and consulting engineering.<br />

He has been employed as a design engineer for a number of manufacturers such as<br />

Warman Pumps (now part of Weir Pumps), Svedala Pumps and Process (now part of<br />

Metso Mineral Systems), Sulzer Pumps North America, and Mazdak Pumps and Mixers.<br />

He has also contracted as a <strong>slurry</strong> and hydraulics s<strong>pe</strong>cialist for major consulting engineering<br />

firms such as ERM, SNC-Lavalin, Fluor, Bateman, Rescan, and Hatch and Associates.<br />

His involvement in the design, expansion, and commissioning of projects has included<br />

ASARCO Ray Tailings (USA), LTV Steel (USA), Zaldivar Pi<strong>pe</strong>line (Chile), Southern<br />

Peru Expansion (Peru), Lomas Bayes (Chile), Escondida (Chile), BHP Diamets (Canada),<br />

Muskeg River Oil Sands (Canada), Bajo Alumbrera (Argentina), Homestake Eskay Creek<br />

(Canada), and many other engineering projects, feasibility studies, and audits.


PREFACE<br />

The science of <strong>slurry</strong> hydraulics started to flourish in the 1950s with simple tests on<br />

pumping sand and coal at moderate concentrations. It has evolved gradually to encompass<br />

the pumping of pastes in the food and process industries, mixtures of coal and oil as a new<br />

fuel, and numerous mixtures of minerals and water. Because of the diversity of minerals<br />

pum<strong>pe</strong>d, the wide range in sizes [43 �m (mesh 325) to 51 mm (2 in)], and the various<br />

physical and chemical pro<strong>pe</strong>rties of the materials, the engineering of <strong>slurry</strong> <strong>systems</strong> requires<br />

various empirical and mathematical models. The engineering of <strong>slurry</strong> <strong>systems</strong> and<br />

the design of pi<strong>pe</strong>lines is therefore fairly complex. This <strong>handbook</strong> targets the practicing<br />

consultant engineer, the maintenance su<strong>pe</strong>rintendent, and the economist. Numerous<br />

solved problems and simplified computer programs have been included to guide the<br />

reader.<br />

The structure of the book is essentially in two parts. The first six chapters form the<br />

first part of the book and focus on the hydraulics of <strong>slurry</strong> <strong>systems</strong>. Chapter 1 is a general<br />

introduction on the preparation of <strong>slurry</strong>, the classification of soils, the siltation of dams,<br />

and the history of <strong>slurry</strong> pi<strong>pe</strong>lines. Chapter 2 focuses on water as a carrier of solids. Chapter<br />

3 progresses with the mechanics of mixing solids and liquids and the principles of rheology.<br />

Chapter 4 presents the various models of heterogeneous flows of settling slurries,<br />

whereas Chapter 5 concentrates on non-Newtonian flows. Due to the importance of o<strong>pe</strong>n<br />

channel flows in the design of long-distance tailings <strong>systems</strong> or <strong>slurry</strong> plants, Chapter 6<br />

was dedicated to a better understanding of these complex flows, which are seldom mentioned<br />

in books on <strong>slurry</strong>. In Part II, the book focuses on components of <strong>slurry</strong> <strong>systems</strong><br />

and their economic as<strong>pe</strong>cts. In Chapter 7, the important equipment of <strong>slurry</strong> processing<br />

plants is presented, including grinding circuits, flotation cells, agitators, mixers, and<br />

thickeners. Chapter 8 presents the guidelines for the design of centrifugal <strong>slurry</strong> pumps,<br />

and methods of correction of their <strong>pe</strong>rformance. Chapter 9 reviews the continuous improvements<br />

of positive displacement <strong>slurry</strong> pumps in their different forms, such as<br />

plunger, diaphragm, or lockhop<strong>pe</strong>r pumps. As <strong>slurry</strong> causes wear and corrosion, as<strong>pe</strong>cts<br />

of the selection of metals and rubbers is presented in Chapter 10. To guide the reader to<br />

the various as<strong>pe</strong>cts of the design of <strong>slurry</strong> pi<strong>pe</strong>lines, Chapter 11 presents practical cases<br />

such as coal, phosphate, limestone, and cop<strong>pe</strong>r concentrate pi<strong>pe</strong>lines. This review of historical<br />

data is followed by a review of standards of the American Society of Mechanical<br />

Engineers and the American Petroleum Institute, as they are extremely useful tools for the<br />

design and monitoring of pi<strong>pe</strong>lines. Finally, as the big unknown is too often cost, Chapter<br />

12 closes the book by offering guidelines for a complete feasibility study for a tailings<br />

disposal system or a <strong>slurry</strong> pi<strong>pe</strong>line.<br />

The author wishes to thank the staff of Mazdak International Inc, particularly Ms.<br />

Mary Edwards for providing typing services with great dedication over a <strong>pe</strong>riod of two<br />

years. The author particularly wishes to thank Fluor Daniel Wright Engineers for allowing<br />

him to use their excellent library in Vancouver, Canada. The author wishes to thank<br />

his former colleagues in a colorful career, particularly Mr. K. Burgess, C.P.Eng. of Warman<br />

International; Mr. A. Majorkwiecz, K. Major, and Mr. Peter Wells of Hatch & Associates;<br />

Mr. I. Hanks, P.Eng. and W. McRae of Bateman Engineering; Mr. R. Burmeister<br />

xvii


xviii PREFACE<br />

H. Basmajian, and Dr. C. Shook, consultants; Mr. C. Hunker, P.Eng, V. Bryant, D.<br />

Bartlett, and W. Li, P.Eng. of Fluor Daniel; and Mr. A. Oak, P.Eng. of AMEC for allowing<br />

him to work on very challenging assignments in Australia and South and North America.<br />

The author wishes to thank the following firms for their contributions in the form of<br />

figures and data to this <strong>handbook</strong>: The Metso Group (formerly the companies Nordberg<br />

and Svedala), Red Valves, Geho Pumps (Weir Pumps), Mobile Pulley and Machine<br />

Works, Inc., Wirth Pumps, Hayward Gordon, Mazdak International Inc., the BHR Group,<br />

and GIW/KSB Pumps.<br />

The author is grateful to the various publishers and associations who allowed him to<br />

reproduce valuable materials in the book.


CHAPTER 8<br />

THE DESIGN<br />

OF CENTRIFUGAL<br />

SLURRY PUMPS<br />

8-0 INTRODUCTION<br />

The centrifugal <strong>slurry</strong> pump is the workhorse of <strong>slurry</strong> flows. Chapter 7 briefed the reader<br />

about some important <strong>slurry</strong> circuits, and it was explained that the grinding circuits consume<br />

a fair portion of the power of a concentrator. One particular pump at the discharge<br />

of the SAG, ball, or other mills is called the mill discharge pump. Wear in these pumps is<br />

particularly harsh, leading to frequent replacement of im<strong>pe</strong>llers and liners, because a fair<br />

portion of the solids remain fairly coarse until recirculated back through the classification<br />

circuit.<br />

The design of centrifugal pumps involves a combination of mathematical and empirical<br />

formulae and models. Although water pumps have been the subject of extensive research<br />

in the past, <strong>slurry</strong> pumps have been designed based on a compromise of what can<br />

be cast with hard alloys, molded in rubber, and what can meet the hydraulic criteria.<br />

A lot of pa<strong>pe</strong>rs have been published over the years on various as<strong>pe</strong>cts of wear in a <strong>slurry</strong><br />

im<strong>pe</strong>ller or volute, <strong>pe</strong>rformance corrections and derating, etc. The reader of these pa<strong>pe</strong>rs<br />

is often left with the impression that the design of these pumps is a combination of science<br />

and art. What is often lacking in the literature are guidelines for the design of <strong>slurry</strong> pumps.<br />

Whereas there are hundreds of manufacturers of water pumps on this planet, the number<br />

of manufacturers of <strong>slurry</strong> and dredge pumps has been reduced to a handful. This<br />

chapter presents some guidelines for the design of <strong>slurry</strong> mill discharge pumps. These<br />

guidelines were develo<strong>pe</strong>d by the author on the basis of the analysis of existing pumps in<br />

the market, throughout his career as a consultant engineer. The designer can vary the<br />

numbers or dimensions presented in the tables of this chapter within a margin of ±15% to<br />

design a pump of his or her choice. These guidelines by themselves must be followed by<br />

pro<strong>pe</strong>r testing, prototy<strong>pe</strong> development, finite element analysis, and ultimately by fieldtesting.<br />

In this chapter, the concepts of ex<strong>pe</strong>ller, pump-out vanes, and dynamic seal will also<br />

be examined. These are very important as<strong>pe</strong>cts of <strong>slurry</strong> pump design that have suffered<br />

from a dearth of information in the published literature.<br />

Wear remains a concern for the design of a <strong>slurry</strong> pump. There is no direct correlation<br />

between the best hydraulics and the highest wear life. In fact, the whole activity of designing<br />

a <strong>slurry</strong> pump is to find an optimum compromise.<br />

8.1


8.2 CHAPTER EIGHT<br />

8-1 THE CENTRIFUGAL SLURRY PUMP<br />

A centrifugal pump is essentially a rotating machine with an im<strong>pe</strong>ller to convert shaft<br />

power into fluid pressure. The dynamic energy is then converted into pressure or head in<br />

a s<strong>pe</strong>cial diffuser or casing. The manufacturers of <strong>slurry</strong> pumps have develo<strong>pe</strong>d a number<br />

of s<strong>pe</strong>cialized designs such as<br />

� Dredge pumps with im<strong>pe</strong>llers as large as 2.6 m (105 in)<br />

� Mill discharge pumps for milling and grinding circuits<br />

� Vertical cantilever pumps (without submerged bearings)<br />

� Froth handling pumps for flotation circuits<br />

� High-pressure tailings and pi<strong>pe</strong>line pumps<br />

� General purpose pumps<br />

� Low-head <strong>slurry</strong> pumps for flue gas desulfurization or flotation circuits<br />

� Submersible <strong>slurry</strong> pumps<br />

The <strong>slurry</strong> pump may be cased in a hard metal (Figure 8-1) or may be cast in iron, with an<br />

internal liner (Figure 8-2), which may be of hard metal or rubber.<br />

The components of the <strong>slurry</strong> pump are divided into two groups:<br />

1. The bearing assembly or cartridge and frame<br />

2. The wetted parts forming the wet end<br />

The main components of the wet end are<br />

� The pump casing volute<br />

� The volute liner<br />

� The front suction plate, or throat bush in large pumps<br />

� The rear wear plate<br />

� The im<strong>pe</strong>ller<br />

� The ex<strong>pe</strong>ller<br />

� The shaft sleeve<br />

� The packing rings<br />

� The stuffing box and gland, greas cup, and associated water connections<br />

� In very s<strong>pe</strong>cial cases the mechanical seal<br />

The drive end of the pump consists of<br />

� The pump shaft<br />

� Piston rings or alternative protection against solids <strong>pe</strong>netrating the bearing assembly<br />

� Forsheda seals or O-rings<br />

� Bearings and bearing nuts<br />

� Grease retaining plates, grease nipples. or oil cup


8.3<br />

adjustment<br />

bolt<br />

bearings<br />

cartridge<br />

shaft sleeve<br />

gland plate<br />

casing joint<br />

stuffing box<br />

water connection<br />

packings rings<br />

frame<br />

back wearplate<br />

im<strong>pe</strong>ller<br />

discharge joint<br />

casing<br />

suction joint<br />

FIGURE 8-1 Components of an unlined hard-metal pump. (Courtesy of Mazdak International Inc.)


8.4 CHAPTER EIGHT<br />

coverplate<br />

coverplate liner<br />

throatbush<br />

suction flange<br />

im<strong>pe</strong>ller<br />

discharge companion flange<br />

backplate liner<br />

backplate<br />

ex<strong>pe</strong>ller<br />

Stuffing box<br />

shaft sleeve<br />

bearing assembly<br />

pump frame<br />

pump shaft<br />

FIGURE 8-2 Components of a rubber-lined <strong>slurry</strong> pump. (Courtesy of Mazdak International<br />

Inc.)<br />

� Bearing cartridge and bearing covers<br />

� An adjustable bolt or mechanism to adjust the im<strong>pe</strong>ller within the casing by moving<br />

the shaft<br />

� The pump frame<br />

� Couplings or pulleys<br />

The purpose of the pump is to produce a certain flow against a certain pressure. This is<br />

done at a certain efficiency. The optimum point at which the efficiency is at a maximum<br />

is called the best efficiency point. For every size or design of pump, there is a best efficiency<br />

point at a given s<strong>pe</strong>ed. The <strong>pe</strong>rformance of the pump is plotted on a curve of head<br />

versus flow (Figures 8-3 and 8-4) By combining different sizes of pumps on a single<br />

chart, a pump tomb chart is produced (Figure 8-5).<br />

Before dwelling on the design of a <strong>slurry</strong> pump, it is essential to have a basic understanding<br />

of the hydraulics involved. But since the design of <strong>slurry</strong> pumps must also take<br />

in account the wear due to pumping abrasive solids, many other factors enter into the<br />

equation, such as the ability to pump large particles and the use of s<strong>pe</strong>cial alloys or polymers<br />

for liners or im<strong>pe</strong>llers.<br />

Practically all <strong>slurry</strong> pumps are single stage. Multistage pumps are limited to mine dewatering<br />

applications. Slurry pumps are rubber lined whenever they are designed to handle<br />

particles finer than 6 mm or 1 – 4�. Because rubber is susceptible to thermal degradation<br />

when the tip s<strong>pe</strong>ed of the im<strong>pe</strong>ller exceeds 28 m/s or 5500 ft/min, rubber-lined pumps are<br />

typically reserved for a maximum head of 30 m (98.5 ft) <strong>pe</strong>r stage.<br />

White iron is a very hard material. It is used in different forms such as Ni-hard and<br />

28% chrome to cast im<strong>pe</strong>llers, casings, and metal liners of <strong>slurry</strong> pumps. Due to concern<br />

about maximum disk stresses, most white iron <strong>slurry</strong> pumps are limited to an im<strong>pe</strong>ller tip<br />

s<strong>pe</strong>ed of 38 m/s or 7480 ft/min. Metal-lined pumps are limited to 55 m or 180 ft <strong>pe</strong>r stage.


Head (ft)<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

Flow rate (L/s)<br />

5<br />

Head vs flow curve<br />

Efficiency Curve<br />

10 15<br />

0 50 100 150 200 250<br />

Flow rate (US gpm)<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

Efficiency (%)<br />

Head (m)<br />

FIGURE 8-3 Performance of a pump showing head versus flow and efficiency versus flow<br />

at constant s<strong>pe</strong>ed.<br />

Head (ft)<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

4200<br />

3900<br />

3600<br />

3300<br />

3000<br />

20%<br />

30%<br />

5<br />

Flow rate L/s<br />

10 15<br />

40%<br />

MINIMUM LIMIT OF USE<br />

45%<br />

48%<br />

2700 r/min<br />

2400<br />

2100<br />

1800<br />

1500<br />

1200<br />

45%<br />

Efficiency curve<br />

0 50 100 150 200 250<br />

Flow rate in US gpm<br />

FIGURE 8-4 Composite curve for the <strong>pe</strong>rformance of a pump showing head versus flow and<br />

efficiency versus flow at various s<strong>pe</strong>eds.<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

Head (m)<br />

best efficiency curve<br />

8.5<br />

s<strong>pe</strong>ed or rotation (rev/min)


8.6 CHAPTER EIGHT<br />

HEAD (METRES)<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0<br />

1105<br />

8X6<br />

816<br />

2000<br />

10X8<br />

667<br />

FLOW IN US GALLONS PER MINUTE<br />

4000<br />

903<br />

575<br />

780<br />

12X10<br />

6000<br />

510<br />

RUBBER RANGE<br />

200 400 600 800 1000 1200<br />

White iron should not be confused with steel. Certain grades of steels are used in <strong>slurry</strong>,<br />

dredging, and phosphate matrix pumps. They are cast at a lower hardness than white<br />

iron and by being more ductile can withstand higher disk stresses. Im<strong>pe</strong>llers cast in steel<br />

can be used in <strong>slurry</strong> pumps up to a tip s<strong>pe</strong>ed of 45 m/s (8858 ft/min).<br />

These are general guidelines, but the consultant engineer should collaborate closely<br />

with the manufacturer. For example, certain s<strong>pe</strong>cial anti-thermal-breakdown additives are<br />

used with some rubbers to exceed the limit of 28 m/s or 5500 ft/min on tip s<strong>pe</strong>ed. In certain<br />

situations, a metal im<strong>pe</strong>ller may be installed with rubber liners, particularly when<br />

there are concerns about <strong>slurry</strong> surges (water hammer) in tailings pi<strong>pe</strong>lines.<br />

8-2 ELEMENTARY HYDRAULICS OF THE<br />

SLURRY PUMP<br />

14X12<br />

8000<br />

691<br />

450<br />

16X14<br />

10000<br />

609<br />

METAL RANGE<br />

390<br />

The correlation between the tip s<strong>pe</strong>ed and the head <strong>pe</strong>r stage is established from basic hydraulics<br />

of im<strong>pe</strong>ller design. There have been two schools in the past for the design of water<br />

pumps—the American school lead by Stepanoff and the Euro<strong>pe</strong>an school lead by Anderson.<br />

The Stepanoff method is based on the concept that an im<strong>pe</strong>ller is designed on the<br />

basis of velocity triangles, and that an ideal volute for best efficiency is then found using<br />

12000<br />

18X16<br />

340<br />

528<br />

14000<br />

FLOW RATE (LITRES/SECONDS)<br />

FIGURE 8-5 “Tomb chart” for pumps showing size of pump versus flow range and head.<br />

20X18<br />

16000<br />

460<br />

18000<br />

20000<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

HEAD (FEET)


various empirical factors. The Anderson school is based on the concept that one of the<br />

most important parameters in pump design is the ratio between the throat area of the volute<br />

and the im<strong>pe</strong>ller discharge area, and therefore more than one volute design can be<br />

matched to a given im<strong>pe</strong>ller.<br />

In the case of <strong>slurry</strong> pumps, passageways are larger than in water pumps to accommodate<br />

solids and the Anderson area ratio is difficult but useful to use. Unfortunately, many<br />

leading references on <strong>slurry</strong> pump design written in North America, such as the work of<br />

Herbrich (1991) and Wilson et al. (1992), continue to ignore the area ratio methods and<br />

focus on the Stepanoff school, which believes that the im<strong>pe</strong>ller is the main producer of<br />

head and efficiency.<br />

The design of a centrifugal <strong>slurry</strong> pump is complex. Performance de<strong>pe</strong>nds on the area<br />

ratio, im<strong>pe</strong>ller tip angle, recirculation patterns, change with wear of the im<strong>pe</strong>ller, back<br />

vanes, and front pump-out vanes.<br />

The flow in an im<strong>pe</strong>ller is fairly complex. A review of the hydraulics is essential to appreciate<br />

wear. In simple terms, a vortex is formed.<br />

8-2-1 Vortex Flow<br />

THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

The vortex creates a pressure field related to the radius from the center of the vortex in<br />

accordance with the following equation:<br />

� = C × R m v0<br />

(8-1)<br />

where<br />

� = angular s<strong>pe</strong>ed of rotating fluids<br />

Rv0 = local radius of vanes<br />

m = exponent<br />

Stepanoff (1993) described various forms of vortices from a free vortex, with angular<br />

velocity � inversely proportional to the square of the radius Rv0, to a su<strong>pe</strong>r-forced<br />

vortex, in which the angular velocity is proportional to the radius, as shown in Table<br />

8-1.<br />

The general distribution of pressure through a vortex, according to Stepanoff, is<br />

� � = C<br />

� � 2 2(m+1)<br />

P Rv0 ��<br />

� + z (8-2)<br />

� 2(m + 1)g<br />

where<br />

C = constant<br />

P = pressure<br />

� = density<br />

m = exponent<br />

g = acceleration due to gravity<br />

z = liquid elevation above the fixed datum<br />

For a forced vortex, the angular s<strong>pe</strong>ed is constant and the liquid revolves as a solid<br />

body. Disregarding friction losses, Stepanoff (1993) claims that no power would be needed<br />

to maintain the vortex. The pressure distribution of this ideal solid body rotation is a<br />

parabolic function of the radius. When the forced vortex is su<strong>pe</strong>rimposed on a radial outflow,<br />

the motion takes the form of a spiral. This is the ty<strong>pe</strong> of flow encountered in a centrifugal<br />

pump. Particles at the <strong>pe</strong>riphery are said to carry the total amount of energy applied<br />

to the liquid.<br />

8.7


8.8 CHAPTER EIGHT<br />

TABLE 8-1 Patterns of Vortex Flow<br />

Angular Peripheral<br />

velocity velocity Pressure<br />

distribution, distribution, distribution,<br />

Case � = C 1 × R m v0 V × R n v0 = C dp = � (�� 2 /g)rdr Ty<strong>pe</strong> of vortex Remarks<br />

1 –� � = C1 × Rv0 � V × Rv0 = C1<br />

2 P/� = C 1 + z1 � = 0, stationary<br />

2 –5/2 � = C2 × Rv0 3/2 V × Rv0 = C2<br />

2 3 P/� = C 2/(3 · g · Rv0) + z2 � is<br />

3 � = C3 × R –2<br />

v0 V × Rv0 = C3 2 P/� = C 3/(2 · g · Rv0) + z3 Z3 + (P/�) + (v2 /(2 · g) higher<br />

4 � = C 4 × R –3/2<br />

v0<br />

5 � = C5 × R –1<br />

6 � = C6 × Rv0 7 � = C 7 × R 0 v0 V × R v0<br />

8 � = C 8 × R 1/2<br />

v0<br />

9 � = C9 × Rv0 V × R –2<br />

10 � = C10 × R m v0 V × Rv0 After Stepanoff (1992).<br />

The parabola shown in Figure 8-6 is a state of equilibrium for a forced vortex and is<br />

similar to a horizontal plane for a stationary fluid. To maintain a flow outward against the<br />

applied pressure, the energy gradient must be smaller than the energy gradient for no<br />

flow. This is what hap<strong>pe</strong>ns in a pump at near shut-off condition, where maximum static<br />

head is obtained without any flow. As flow increases through the im<strong>pe</strong>ller, the head<br />

drops.<br />

In the case of the ex<strong>pe</strong>ller, the designer tries to reach the parabola for energy gradient<br />

without flow. However, as Case 7 in Table 8-1 shows, the pressure gradient is a square<br />

function of R and inversely proportional to the square of the angular velocity. And in fact,<br />

below a certain angular velocity, there is not enough pressure to overcome the difference<br />

between volute and outside atmospheric pressure. The ex<strong>pe</strong>ller or dynamic seal then stops<br />

<strong>pe</strong>rforming and leakage occurs.<br />

8-2-2 The Ideal Euler Head<br />

= constant, free vortex toward<br />

V × R 1/2<br />

v0 = C 4 P/� = –C 4/(g · r) + z 4 center of<br />

v0 V × R0 v0 = C5 2 P/� = [C 5/g] · log Rv0 + z5 V = constant the vortex<br />

–1/2 –1/2 V × Rv0 = C6<br />

2 P/� = C 6 · Rv0/g + z6 V2 /Rv0 = constant<br />

–1 = C7<br />

2 2 P/� = C 7 · Rv0/(2 · g) + h7 = centrifugal force<br />

� = constant, � =<br />

forced vortex constant<br />

–3/2 V × Rv0 = C8 P/� = C82 · R3 v0/(3 · g) + h8 Su<strong>pe</strong>r forced vortex � is<br />

v0 = C9 P/� = C9 · R 4 v0/(4 · g) + h9 Su<strong>pe</strong>r forced vortex higher<br />

–(m+1) = C10 P/� = [C 2 2(m+1) Rv0 ]/ General form of su<strong>pe</strong>r toward<br />

[2(m + 1) · g] + h forced vortex <strong>pe</strong>riphery<br />

of vortex<br />

The ideal pressure that a pump im<strong>pe</strong>ller can develop is called the Euler pressure. Consider<br />

the flow through a radial im<strong>pe</strong>ller between two radii R1 and R2. The im<strong>pe</strong>ller is rotating<br />

at an angular s<strong>pe</strong>ed � (in rad/s) so that the <strong>pe</strong>ripheral s<strong>pe</strong>eds are res<strong>pe</strong>ctively:<br />

U1 = R1 · � (8-3a)<br />

U2 = R2 · � (8-3b)<br />

The liquid flows radially at a meridional velocity Cm, <strong>pe</strong>r<strong>pe</strong>ndicular to the <strong>pe</strong>ripheral<br />

velocity U. The value of Cm is determined from continuity equation, It is necessary to take<br />

into account the local area of the flow, which is a function of the radius and the width of


THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

FIGURE 8-6 Pressure distribution in an im<strong>pe</strong>ller versus radius for condition of flow and no<br />

flow. (From Stepanoff, 1993. Reprinted by <strong>pe</strong>rmission of Krieger Publishers.)<br />

the channel, minus the blockage area due to the finite thickness and angle of inclination of<br />

the blades.<br />

The channels between the im<strong>pe</strong>ller vanes follow a certain profile. At the intersection<br />

with the radius under consideration, the angle between the vane and the tangent to the radius<br />

is defined as �. A component of velocity is in the direction of � and is called the relative<br />

velocity W.<br />

The vector addition of U and W result in the absolute velocity V. Both V and W share<br />

the same component of meridional s<strong>pe</strong>ed Cm; a vector representation is shown in Figure<br />

8-8.<br />

The Euler “total” head between radii R1 and R2 is defined as<br />

HE = (8-4)<br />

where<br />

(V 2 – V 2 1) = change in absolute kinetic energy<br />

(W 2 2 – W 2 1) = change in relative kinetic energy<br />

(U 2 2 – U 2 1) + (W 2 2 – W 2 (V<br />

1) = change in static energy through the im<strong>pe</strong>ller<br />

2 – V 2 1) – (U 2 2 – U 2 1) + (W 2 2 – W 2 1)<br />

����<br />

2g<br />

It is clear that W = C m · cot �. Static head rise is<br />

gH s = (U 2 2 – U 2 1) + (C m2 · cot � 2) 2 – (C m1 · cot � 1) 2 (8-5)<br />

Furthermore because the curvature of the front and back shrouds of an im<strong>pe</strong>ller, are different,<br />

the meridional velocity is not uniform and may be higher toward the back shroud.<br />

For a linear variation of the meridional velocity between the front and back shrouds (Figure<br />

8-7), Stepanoff (1993) derived the following equation for theoretical head:<br />

U 2 2<br />

� g<br />

U2Cm2 �<br />

g tan �2<br />

(V2 – V1) 2<br />

��<br />

H t = � � – � 1 + � (8-6)<br />

12 Cm 2<br />

8.9


8.10 CHAPTER EIGHT<br />

FIGURE 8-7 Pressure and velocity distribution for cases shown in Table 8-1. (From<br />

Stepanoff, 1993. Reprinted by <strong>pe</strong>rmission of Krieger Publishers.)<br />

The term 1 + [(V 2 – V 1) 2 /12 C m 2 ] is greater for the wide im<strong>pe</strong>llers encountered in mining<br />

<strong>slurry</strong> pumps. For <strong>slurry</strong> pumps, the value of � 1 at the tip diameter of the eye of the<br />

im<strong>pe</strong>ller is between 14 and 30 degrees. The value of � 2 at the tip diameter of the vanes is<br />

typically between 25 and 35 degrees. Stepanoff (1993) has indicated that inlet angles as<br />

high as 50 degrees are used on water pumps. This is, however, not the case with <strong>slurry</strong><br />

pumps, as prerotation causes tremendous wear of the throat bush.<br />

The vast majority of modern pumps have a discharge angle � 2 smaller than 90 degrees.<br />

They are called im<strong>pe</strong>llers and have backward curved vanes. Ex<strong>pe</strong>llers are often designed<br />

with radial vanes (i.e., � 2 = 90 degrees). Forward vanes with � 2 larger than 90 de-<br />

W 1<br />

C m1<br />

1 1<br />

U1 V1<br />

Inlet velocities at R 1<br />

rotation<br />

W 1<br />

C<br />

m1<br />

1<br />

1<br />

V<br />

U 1<br />

1<br />

R 2<br />

R 1<br />

2<br />

W2<br />

FIGURE 8-8 Ideal velocity profile in an im<strong>pe</strong>ller.<br />

U 2<br />

Outlet velocities at R 2<br />

m2<br />

W 2 C<br />

2<br />

V<br />

2<br />

V<br />

C 2<br />

m2<br />

2<br />

2<br />

U<br />

2


THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

grees are restricted to very low flow and high-head pumps and to some ex<strong>pe</strong>llers. Theoretically,<br />

an im<strong>pe</strong>ller with forward vanes would give a higher static head rise. Unfortunately,<br />

it is also the largest consumer of power and is considered to be inefficient.<br />

Clay and other slurries can be very viscous. Herbrich (1991) has suggested using discharge<br />

angle � 2 as high as 60 degrees on im<strong>pe</strong>llers for very viscous slurries but did not<br />

produce data to support such a suggestion. Stepanoff (1993) recommended the following<br />

design procedures for s<strong>pe</strong>cial pumps. These pumps would be suited to pump viscous liquids,<br />

but their <strong>pe</strong>rformance may be impaired on water.<br />

1. Use high im<strong>pe</strong>ller discharge angles up to 60 degrees to reduce the im<strong>pe</strong>ller diameter<br />

necessary to produce the same head and effectively reduce disk friction losses. Consequently,<br />

the im<strong>pe</strong>ller channels become shorter and the im<strong>pe</strong>ller hydraulic friction is reduced.<br />

2. Eliminate close-clearance wide sealing rings at the im<strong>pe</strong>ller eye and provide knifeedge<br />

seals (one or two) similar to those used on blowers. Leakage loss becomes secondary<br />

when pumping viscous liquids.<br />

3. Provide a similar axial seal at the im<strong>pe</strong>ller outside diameter to confine the liquid between<br />

the im<strong>pe</strong>ller and casing walls. This in turns raises the tem<strong>pe</strong>rature of the liquid<br />

in the confined space (due to friction) well above the tem<strong>pe</strong>rature of the remaining liquid<br />

passing through the im<strong>pe</strong>ller. Due to the tem<strong>pe</strong>rature effects, viscosity is artificially<br />

reduced and disk friction losses are trimmed down. In fact, Stepanoff (1993) goes as<br />

far as suggesting injecting a light or heated oil in the confined space to reduce power<br />

loss due to friction.<br />

4. Provide an ample gap (twice the normal) between the casing tongue or cutwater and<br />

the im<strong>pe</strong>ller outside diameter. Otherwise, the shrouds of the im<strong>pe</strong>ller would produce<br />

head by viscous drag at low capacities, and would decrease the efficiency of pumping.<br />

5. High rotational s<strong>pe</strong>ed and high s<strong>pe</strong>cific s<strong>pe</strong>ed lead to better efficiency and more head<br />

capacity output than low s<strong>pe</strong>cific s<strong>pe</strong>ed pumps on viscous liquids.<br />

These recommendations were written with very viscous fluids in mind. Obviously,<br />

points 2 and 3 would not apply to a <strong>slurry</strong> pump. However, <strong>slurry</strong> pumps may use pumpout<br />

vanes, which effectively are dynamic seals. These recommendations can be modified<br />

to suit the design of s<strong>pe</strong>cial pumps for viscous slurries. The field of <strong>slurry</strong> pumps for very<br />

viscous slurries and difficult flotation frothy <strong>slurry</strong> associated with the oil sands industry<br />

is continuously evolving.<br />

In some cases of pumping oil sand froth, it has been found that injecting 1% of water<br />

or a light oil as a lubricant just at the suction of the pump can improve the efficiency of<br />

the pump.<br />

8-2-3 Slip of Flow Through Im<strong>pe</strong>ller Channels<br />

8.11<br />

Due to the curvature of the vane, the flow on the up<strong>pe</strong>r surface of a vane is usually faster<br />

than the flow on the lower surface of the vane. If we consider the direction of rotation, the<br />

up<strong>pe</strong>r surface is also called the advancing surface or leading surface. The lower surface is<br />

the trailing surface. The pressure being higher on the trailing surface, the fluid leaves tangentially<br />

only at the trailing surface. A certain amount of liquid is attracted by the lower<br />

surface of the following vane and a pattern of flow recirculation develops as shown in<br />

Figure 8-9. To compare this situation with that of an airplane, which many of us have examined,<br />

vortices form behind a flying wing, as air tends to roll from the up<strong>pe</strong>r pressure


8.12 CHAPTER EIGHT<br />

R<br />

2 R 1<br />

zone of the lower surface toward the lower pressure zone above the wing. A vortex sheet,<br />

called “horseshoe vortex,” forms behind the airplane wing.<br />

The velocities in a real im<strong>pe</strong>ller do not follow the ideal “Euler” im<strong>pe</strong>ller pattern, and a<br />

degree of “slip” occurs. The angles of flow and forces deviate from the theoretical values<br />

as shown in Figure 8-10 by a “lag” angle. The slip factor is in fact as the ratio of the measured<br />

absolute velocity to the theoretical Euler absolute velocity at the tip diameter of the<br />

vanes:<br />

� = V2�/V2 (8-7)<br />

Since the average meridional velocity is essentially a function of the ratio of flow rate<br />

to the discharge area at the tip of the im<strong>pe</strong>ller, it is not affected by slip. However, a change<br />

in the absolute velocity is accompanied by changes in the relative velocity and of the angles<br />

with res<strong>pe</strong>ct to the <strong>pe</strong>ripheral tangential s<strong>pe</strong>ed. Various equations have been develo<strong>pe</strong>d<br />

over the years to evaluate the slip factor. The most famous is Stodola’s formula:<br />

� · sin �2 � = 1 – ���� (8-8)<br />

Z<br />

where Z = number of vanes. Stodola’s formula was originally develo<strong>pe</strong>d for zero flow, but<br />

has been extensively used for design flows of water pumps even at best efficiency point.<br />

Another equation used to determine slip was develo<strong>pe</strong>d by Pfeiderer (1961):<br />

a<br />

�<br />

Z<br />

�2 �<br />

60<br />

rotation of im<strong>pe</strong>ller<br />

relative recirculation<br />

FIGURE 8-9 Recirculation in pump im<strong>pe</strong>llers (after Stepanoff, 1993).<br />

R 2 2<br />

� S<br />

� = 1/� 1 + � 1 + � � (8-9)<br />

theoretical V' 2 measured<br />

W W'<br />

2 2 V2<br />

2<br />

2<br />

U 2<br />

2<br />

2<br />

(with slip)<br />

FIGURE 8-10 Slip of flow in im<strong>pe</strong>llers versus ideal velocity profile.


where<br />

S = � R2 R<br />

R dR (8-10)<br />

S is called the static moment and is obtained by graphical integration along the meridional<br />

plane of the vanes. In the s<strong>pe</strong>cial case of a cylindrical vane<br />

S = � R2 R<br />

R dR = |(R 2 2 – R 1 2 )<br />

and the slip factor is<br />

� = 1/�1 + �1 + a �2 2<br />

� ����� (8-11)<br />

2 2 Z 60 1 – (R 1/R 2) In the s<strong>pe</strong>cial case in which R1/R2 is smaller than 0.5, the slip does not increase anymore,<br />

and a ratio R1/R2 = 0.5 should be assumed.<br />

The magnitude “a” de<strong>pe</strong>nds on the design of the casing. Pfeiderer (1961) established<br />

the following values for the coefficient “a”:<br />

Volute a = 0.65–0.85<br />

Vaned diffuser a = 0.60<br />

Vaneless diffuser a = 0.85 – 1.0<br />

THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

8.13<br />

Most <strong>slurry</strong> pumps use a volute (Figure 8-11). Vaned diffusers are used in certain<br />

mine dewatering pumps.<br />

Defining the hydraulic efficiency as �H, the head develo<strong>pe</strong>d by the pump is expressed<br />

as:<br />

H = ��HU2V2 (8-12)<br />

Equation (8-12) establishes the effect of the casing and the im<strong>pe</strong>ller on the head develo<strong>pe</strong>d<br />

at the so-called best efficiency point. Because of the rather simplistic Stodala equa-<br />

volute casing diffuser vane casing<br />

FIGURE 8-11 Volute and vaned diffusers.


8.14 CHAPTER EIGHT<br />

TABLE 8-2 Test Data from Herbich (1991)<br />

Velocity Theoretical Measured<br />

Radial 4.21 ft/s (1.3 m/s) 15.6 ft/s (4.8 m/s)<br />

Tangential 55.80 ft/s (17.0 m/s) 39.6 ft/s (5.2 m/s)<br />

tion (8-8), it is sometimes assumed that the im<strong>pe</strong>ller is the main contributor to head. The<br />

equation for the head is also expressed in terms of the discharge angle from the vanes, the<br />

slip factor, and the hydraulic efficiency as:<br />

H = ��H �1 – � Cm2 · cot<br />

�2 ��<br />

(8-13)<br />

U2<br />

Herbich (1991) measured extensively the lag angle and deviation from theoretical angles<br />

in the case of the Essayon dredge pump and reported two cases of im<strong>pe</strong>ller tip vane discharge<br />

angle �2 (Table 8-2). In the first case, the vane was designed with a physical tip angle<br />

at the vanes of 22.5°. This would have been theoretically the angle for the relative s<strong>pe</strong>ed<br />

W. However, test data measured an average angle of 30.5°. In the second case, the vane had<br />

a discharge angle of 35° but test data indicated that the relative velocity was effectively inclined<br />

at an average angle of 12°. In fact there is no definite value. In the case of the first<br />

im<strong>pe</strong>ller with a vane angle of 22.5°, at a flow rate of 63 L/s (1000 gpm) the flow between<br />

the channels was measured to have streams inclined between 61° on the lower surface and<br />

25° on the forward surface with various values between 21 47°. A different pattern was noticed<br />

at 38 L/s (600 gpm). The distortion of the flow is therefore a function of the ratio of<br />

flow rate to normalized flow (at best efficiency point). When the ex<strong>pe</strong>rimental angle is<br />

higher than the theoretical, Herbich pointed out that it would mean that the particles tend to<br />

avoid contact, thus minimizing the possibility of scour. On the other hand, if the measured<br />

angle is less than the theoretical, the solids will impact the vanes and cause wear.<br />

Because it is difficult to measure slip, an ex<strong>pe</strong>rimental head coefficient is defined as:<br />

�SI = � 2g<br />

U<br />

HBEP � (8-14)<br />

2 U 2<br />

For some historical reasons, U.S. books drop the numerator 2:<br />

2 2<br />

�<br />

g<br />

�US = � (8-15)<br />

2 U 2<br />

The reader must therefore be careful when comparing pumps manufactured in North<br />

America with those manufactured in Euro<strong>pe</strong>.<br />

8-2-4 S<strong>pe</strong>cific S<strong>pe</strong>ed<br />

gH BEP<br />

The steepness of the curve between the best efficiency capacity and the shut-off point of<br />

the pump de<strong>pe</strong>nds on the geometrical design of im<strong>pe</strong>ller and casing. With so many different<br />

designs of pumps, engineers have used nondimensional s<strong>pe</strong>cific s<strong>pe</strong>eds and other parameters.<br />

In the International System of Units, the s<strong>pe</strong>cific s<strong>pe</strong>ed is defined as:<br />

N · �Q�<br />

Nq = � (8-16)<br />

3/4 H


THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

where<br />

N = rotational s<strong>pe</strong>ed in rev/min<br />

Q = capacity in cubic meters <strong>pe</strong>r second, at best efficiency capacity<br />

H = differential head in meters at best efficiency capacity<br />

The s<strong>pe</strong>cific s<strong>pe</strong>ed in the United States is defined as:<br />

N · �Q�<br />

NUS = � (8-17)<br />

3/4 H<br />

where<br />

N = rotational s<strong>pe</strong>ed in rev/min<br />

Q = capacity in U.S. gallons <strong>pe</strong>r minute, at best efficiency capacity<br />

H = differential head in feet at best efficiency capacity<br />

Some books include the acceleration of gravity g or 32.2 ft/s in the denominator for<br />

the sake of consistency, but for historical reasons Equation 8-17 is used.<br />

Another term sometimes used in international books is the characteristic number:<br />

� · �Q�<br />

Ks = � (8-18)<br />

3/4 [gH]<br />

Most <strong>slurry</strong> pumps o<strong>pe</strong>rate at a s<strong>pe</strong>cific s<strong>pe</strong>ed smaller than 2000 in U.S. units or 39 in<br />

SI units. In this range, the tip diameter of the im<strong>pe</strong>ller may be between 2 to 3.5 folds of<br />

the suction diameter. The shut-off head is then between 150% and 110% of the best efficiency<br />

point head at the same s<strong>pe</strong>ed (Figure 8-12).<br />

Addie and Helmly (1989) measured the head coefficient (as defined in the United<br />

States) and the efficiency of a number of <strong>slurry</strong> and dredge pumps. Their results are<br />

shown in Figures 8-13 and 8-14. They pointed out that the <strong>slurry</strong> and dredge pumps were<br />

on the average between 5% and 9% less efficient than their water counterparts.<br />

Example 8-1<br />

A <strong>slurry</strong> pump is to be designed for a head at best efficiency of 150 ft at a flow rate of<br />

1200 gpm. Assuming a head coefficient of 0.5 (by U.S. definition), determine the diameter<br />

and the s<strong>pe</strong>ed of rotation if the s<strong>pe</strong>cific s<strong>pe</strong>ed is 1100 (in U.S. units).<br />

Solution in USCS Units<br />

From Equation 8-15:<br />

gHBEP 32.2 × 150<br />

�US = � = 0.5 = ��<br />

2 2 U 2<br />

U 2<br />

U2 = 98.3 ft/s<br />

From Equation 8-17, the s<strong>pe</strong>cific s<strong>pe</strong>ed in the United States is defined as:<br />

Ns = N · Q1/2 /H 3/4 = 1100 = N · 12001/2 /1503/4 8.15<br />

N = 889 rpm = 93.1 rad/sec<br />

Since U = R�, then R = 98.3/93.1 = 1.06 ft. The im<strong>pe</strong>ller diameter is therefore 2.11 ft or<br />

25.3 inch (643 mm).<br />

Every manufacturer has their proprietary design criteria, and for a given size some<br />

manufacturers may have an im<strong>pe</strong>ller design that pumps more than others. In the case of


8.16<br />

FIGURE 8-12 Sha<strong>pe</strong> of im<strong>pe</strong>ller versus s<strong>pe</strong>cific s<strong>pe</strong>ed in USCS units. [From I. Karasik et al. (Eds.), Pump Handbook, reprinted by <strong>pe</strong>rmission from<br />

McGraw-Hill.]


THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

8.17<br />

FIGURE 8-13 Head coefficient versus s<strong>pe</strong>cific s<strong>pe</strong>ed from Addie and Helmly (1989) (reproduced<br />

by <strong>pe</strong>rmission of Central Dredging Association, Delft, Netherlands). This plot is somewhat<br />

confusing as it uses the U.S. definition of the head coefficient (as <strong>pe</strong>r Equation 8-15)<br />

against the s<strong>pe</strong>cific s<strong>pe</strong>ed in SI units. The reader should multiply the head coefficient by a factor<br />

of 2 to use the SI definition of head coefficient as <strong>pe</strong>r Equation 8-14.<br />

<strong>slurry</strong> pumps, attention must be paid to the wear life of the pump. Too little flow in a large<br />

pump leads to excessive recirculation, and too much flow would cause rapid wear. The<br />

relationship between the volute sha<strong>pe</strong> and the im<strong>pe</strong>ller plays a major role, too. These parameters<br />

are refined through detailed engineering and field-testing. A good starting point<br />

for the design of mill discharge pumps is shown in Tables 8-3 and 8-4. These are realistic<br />

values that mills ex<strong>pe</strong>ct from pumps.<br />

The next step is to define the steepness of the curve. Slurry pumps are designed to be<br />

forgiving as processes too often change. Very steep curves are not encouraged, but flat<br />

curves do create overloading problems to the drivers. A shut-off head in the range of<br />

125% to 135% of the best efficiency head is recommended. The <strong>slurry</strong> pump design engineer<br />

should then establish what is often referred to as a 5-points curve, as shown in Tables<br />

8-5 and 8-6.and Figure 8-15.<br />

As early as 1938, Anderson develo<strong>pe</strong>d a concept of the ratio of the area of flows between<br />

the vanes of the im<strong>pe</strong>ller and the throat area (Figure 8-10) that is basic to the <strong>pe</strong>rformance<br />

of the pump. His methodology is called the “area ratio” (Figure 8-16). Worster<br />

(1963) demonstrated this to be correct by mathematical derivation. Anderson (1977,<br />

1980, 1984) extended his analysis by statistical analysis of a large number of water pumps<br />

and turbines. Unfortunately, no similar work has been done on <strong>slurry</strong> pumps and because<br />

<strong>slurry</strong> im<strong>pe</strong>llers are fairly wider than water pump im<strong>pe</strong>llers to allow the passage of rocks<br />

and large particles, the Worster curves do not apply well to the design of solids-handling<br />

pumps.<br />

Not all applications of pump slurries require wide im<strong>pe</strong>llers. In fact in the last 20<br />

years, grinding circuits have greatly evolved to the point that very fine ores are pum<strong>pe</strong>d.<br />

For these applications, narrower and more efficient im<strong>pe</strong>llers should designed.


8.18 CHAPTER EIGHT<br />

FIGURE 8-14 Efficiency of large dredge pumps versus s<strong>pe</strong>cific s<strong>pe</strong>ed (in SI units). (From Addie<br />

and Helmly, 1989. Reproduced by <strong>pe</strong>rmission of Central Dredging Association, Delft, Netherlands)<br />

8-2-5 Net Positive Suction Head and Cavitation<br />

When the pressure on the suction flange of the pump is insufficient, the pump starts to<br />

cavitate and becomes very noisy. The net positive suction head (NPSH; see Figure 8-17)<br />

is the absolute head above the vapor pressure at the suction flange of the pump:<br />

P 2<br />

e – PD – PV V e<br />

NPSHA = �� + Z1 – Z2 � (8-19)<br />

�g<br />

�g<br />

where<br />

Pe = pressure at the surface of the liquid in absolute terms on the suction side<br />

PD = pressure losses between the surface of the liquid and the pump, due to friction,<br />

valves, etc.<br />

PV = vapor pressure<br />

Z1 = geodetic elevation of liquid surface above the centerline of the pump im<strong>pe</strong>ller<br />

Ze= geodetic elevation of the centerline of the pump im<strong>pe</strong>ller<br />

Ve = suction s<strong>pe</strong>ed at the eye of the im<strong>pe</strong>ller


THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

TABLE 8-3 Recommendations for Design of Rubber-Lined Mill Discharge Pumps<br />

TABLE 8-4 Recommendations for Design of Metal-Lined or Hard Metal Mill<br />

Discharge Pumps<br />

8.19<br />

Size<br />

suction to<br />

discharge, inch<br />

Flow<br />

_____________<br />

Head<br />

___________ Efficiency<br />

Suction s<strong>pe</strong>ed<br />

____________<br />

Discharge s<strong>pe</strong>ed<br />

_____________<br />

(mm/mm) L/s US gpm m ft % m/s ft/s m/s ft/s<br />

8 × 6<br />

200/150<br />

130 2061 30 98.5 70 4.2 13.7 7.2 23.5<br />

10 × 8<br />

250 × 200<br />

220 3487 30 98.5 74 4.5 14.8 6.7 22.1<br />

12 × 10<br />

300 × 250<br />

310 4915 30 98.5 76 4.4 14.4 6.12 20.1<br />

14 × 12<br />

350/300<br />

425 6737 30 98.5 79 4.4 14.5 5.86 19.2<br />

16 × 14<br />

400 × 300<br />

560 8877 30 98.5 81 4.3 14.1 5.64 18.5<br />

18 × 16<br />

450 × 400<br />

685 10859 30 98.5 83 4.3 14.1 5.45 17.9<br />

20 × 18<br />

500/450<br />

875 13870 30 98.5 84 4.3 14.2 5.33 17.5<br />

From Abulnaga (2001). Courtesy of Mazdak International Inc.<br />

Size<br />

suction to<br />

Flow<br />

discharge, inch _____________<br />

Head<br />

___________ Efficiency<br />

Suction s<strong>pe</strong>ed<br />

____________<br />

Discharge s<strong>pe</strong>ed<br />

_____________<br />

(mm/mm) L/s US gpm m ft % m/s ft/s m/s ft/s<br />

8 × 6<br />

200/150<br />

176 2797 55 180 70 5.7 18.6 9.7 32<br />

10 × 8<br />

250 × 200<br />

298 4732 55 180 74 6.1 20 9.1 30<br />

12 × 10<br />

300/250<br />

421 6670 55 180 76 6 19.5 8.3 27<br />

14 × 12<br />

350/300<br />

577 9143 55 180 79 6 19.7 8 27.3<br />

16 × 14<br />

400/300<br />

760 12047 55 180 81 5.8 19.3 8 25.1<br />

18 × 16<br />

450/400<br />

924 14647 55 180 83 5.8 19.3 7.4 24.1<br />

20 × 18<br />

500/450<br />

1188 18823 55 180 84 5.8 19.3 7.2 23.8<br />

From Abulnaga (2001). Courtesy of Mazdak International Inc.


8.20 CHAPTER EIGHT<br />

TABLE 8-5 Preliminary Range for Efficiency versus Flow (L/s units) For Mill Discharge<br />

Pump—Rubber-Lined Version<br />

Pump Size (suction/discharge)<br />

8 × 6 in 10 × 8 in 12 × 10 in 14 × 12 in 16 × 14 in 18 × 16 in 20 × 18 in<br />

Ratio of<br />

flows,<br />

Ratio of<br />

efficiency,<br />

200 × 150 250 × 200 300 × 250 350 × 300 400 × 350 450 × 400 500 × 450<br />

mm mm mm mm mm mm mm<br />

Q/QN �/�BEP Flow in L/s<br />

0.25 0.5 32.5 55 77.5 106 140 171 219<br />

0.5 0.8 65 110 155 213 280 342 438<br />

0.75 0.95 97.5 165 232.5 319 420 523 656<br />

1.00 1.0 130 220 310 425 560 684 875<br />

1.15 0.97 150 253 356.5 489 644 787 1006<br />

From Abulnaga (2001). Courtesy of Mazdak International Inc.<br />

Each pump has a minimum required NPSH that is established through testing. It is defined<br />

as the required NPSH or NPSHR. The suction-s<strong>pe</strong>cific s<strong>pe</strong>ed is defined at the best<br />

efficiency point as:<br />

N · �Q�<br />

NSS = ��<br />

(8-19)<br />

NPSHR 3/4<br />

The value of NPSH is established at the point where the suction conditions at best efficiency<br />

flow suffer a 3% drop of total dynamic head.<br />

Solids present in <strong>slurry</strong> do not contribute to the vapor pressure, but they contribute to<br />

the density of the mixture as well as to the friction or pressure losses on the suction. This<br />

could be confusing to the inex<strong>pe</strong>rienced engineer who has to handle water vapor pressure<br />

as well as <strong>slurry</strong> density. One approach is to calculate the pressure on the suction in units<br />

of pressure and then to convert back into units of length.<br />

TABLE 8-6 Preliminary Range for Efficiency versus Flow (L/s units), Metal-Lined or<br />

Hard Metal Mill Discharge Pumps<br />

Pump Size (suction/discharge)<br />

8 × 6 in 10 × 8 in 12 × 10 in 14 × 12 in 16 × 14 in 18 × 16 in 20 × 18 in<br />

Ratio of<br />

flows,<br />

Ratio of<br />

efficiency,<br />

200 × 150 250 × 200 300 × 250 350 × 300 400 × 350 450 × 400 500 × 450<br />

mm mm mm mm mm mm mm<br />

Q/QN �/�BEP Flow in L/s<br />

0.25 0.5 44 74.5 105 144 190 231 297<br />

0.5 0.8 88 149 210 288.5 280 462 594<br />

0.75 0.95 132 223.5 316 315.8 420 693 891<br />

1.00 1.0 176 298 421 577 760 924 1188<br />

1.15 0.97 202 342.7 484 664 874 1063 1366<br />

From Abulnaga (2001). Courtesy of Mazdak International Inc.


H/H N<br />

1.2<br />

0.0<br />

0.0<br />

THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

HEAD<br />

1.0 1.0<br />

0.8 0.8<br />

0.6 0.6<br />

EFFICIENCY<br />

0.4 0.4<br />

0.2 0.2<br />

0.5 1.0<br />

0.0<br />

1.5<br />

Q/Q<br />

N<br />

It is often recommended that the available NPSH be at least 0.9 m or approximately 3<br />

ft higher than the required NPSH as shown on the pump curve.<br />

Example 8-2<br />

Slurry with a s<strong>pe</strong>cific gravity of 1.48 is to be pum<strong>pe</strong>d from a pond 3 m lower than the centerline<br />

of the im<strong>pe</strong>ller. The pond is situated at a high altitude. The atmospheric pressure is<br />

85 kPa. The friction losses have been determined to be 1.5 m. The vapor pressure of water<br />

is 4.24 kPa. The <strong>slurry</strong> enters the pump at a velocity of 3.5 m/s. Determine the available<br />

NPSH.<br />

Solution<br />

Pressure due to friction losses is:<br />

�gH = 1480 · 9.81 · 1.5 = 21,778 Pa<br />

The geodetic elevation of the centerline of the pump im<strong>pe</strong>ller is 3 m higher than the liquid;<br />

this results in a negative pressure or<br />

�g�Z = 1480 · 9.81 · (–3) = –43,556 Pa<br />

Dynamic head losses due to a velocity of 3.5 m/s are:<br />

1480 · 3.52 /2 = 9065 Pa<br />

Net positive pressure is:<br />

85,000 – 43,556 – 21,778 – 9065 – 4240 = 24,491 Pa<br />

Converting back into head of water:<br />

24,491/(9.81 · 1000) = 2.496 m of water<br />

N<br />

8.21<br />

FIGURE 8-15 Normalized curves of head and efficiency versus values at the best efficiency<br />

point.


8.22 CHAPTER EIGHT<br />

FIGURE 8-16 The area ratio curves for water pumps. No similar curves have been published<br />

for <strong>slurry</strong> or dredge pumps. (From Worster, 1963. Reproduced by <strong>pe</strong>rmission of the Institution<br />

of Mechanical Engineers, UK.<br />

This is very low, and since the engineer must avoid cavitations, he or she may consider<br />

the use of a submersible <strong>slurry</strong> pump or a vertical cantilever pump instead of a horizontal<br />

pump on the shore.<br />

The NPSH can be expressed as the function of suction s<strong>pe</strong>ed and the eye tip s<strong>pe</strong>ed at<br />

the suction diameter (Turton, 1994):<br />

2 2 0.9 C m + 0.115 U 1<br />

NPSH = ��<br />

(8-20)<br />

9.81


8.23<br />

Z S<br />

Liquid at<br />

vapor pressure P v<br />

Z<br />

E<br />

Absolute Atmospheric Pressure<br />

P<br />

A<br />

Pressurized gas at surface at<br />

gauge pressure P<br />

B<br />

H<br />

1<br />

FIGURE 8-17 Concept of net positive suction head.<br />

P = P + P<br />

E A B<br />

Pressure due to<br />

friction losses<br />

P<br />

D


8.24 CHAPTER EIGHT<br />

Example 8-3<br />

A pump im<strong>pe</strong>ller rotates at 500 rpm to pump 65 L/s through a suction diameter of 200<br />

mm. Using Equation 8-20, determine the required NPSH.<br />

Solution<br />

The velocity Cm is determined by dividing the flow rate by the suction area:<br />

Cm = 0.065/[0.25 · � · 0.22 ] = 2.07 m/s<br />

U = 2�RN/60 = 2 · � · 0.1 · 500/60 = 5.24 m/s<br />

0.9 · 2.07<br />

NPSH = = 0.715 m<br />

2 + 0.115 · 5.242 ���<br />

9.81<br />

In reality, NPSH de<strong>pe</strong>nds on many other factors, particularly clearances at the im<strong>pe</strong>ller<br />

eye, prerotation, the use of inducers, etc. Many empirical studies tend to support<br />

that a low NPSH im<strong>pe</strong>ller should have a vane entry angle of 14° to 15°.<br />

A cavitations parameter � is defined as the ratio of required NPSH to the pump total<br />

dynamic head at the best efficiency point at the given s<strong>pe</strong>ed:<br />

NPSH<br />

� = � (8-21)<br />

TDH<br />

Addie and Helmly (1989) measured the cavitations parameter against s<strong>pe</strong>cific s<strong>pe</strong>ed<br />

for a number of dredging pumps. Their work is represented in Figure 8-18. Tables 8-7 and<br />

8-8 also show certain calculations for the design of mill discharge pumps.<br />

FIGURE 8-18 Cavitation factor versus s<strong>pe</strong>cific s<strong>pe</strong>ed (in metric units) for <strong>slurry</strong> and dredge<br />

pumps. (From Addie and Helmly, 1989. Reproduced by <strong>pe</strong>rmission of Central Dredging Association,<br />

Delft, Netherlands)


8-3 THE PUMP CASING<br />

THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

8.25<br />

TABLE 8-7 Recommendations for Im<strong>pe</strong>ller Diameter, S<strong>pe</strong>ed, S<strong>pe</strong>cific S<strong>pe</strong>ed Number,<br />

and Cavitations Parameter for Rubber-Lined Mill Discharge Pumps (U.S. Units)<br />

Head Sigma<br />

Vane d2, d2/dS, S<strong>pe</strong>ed, Flow, Efficiency, Head, S<strong>pe</strong>cific factor, cavitation<br />

Model inch tip/suction rpm US gpm % ft s<strong>pe</strong>ed �US factor<br />

8 × 6 in 20.87 3.48 816 2061 70 98.5 1186 0.14 0.14<br />

10 × 8 in 26.77 3.35 667 3487 74 98.5 1261 0.13 0.16<br />

12 × 10 in 31.10 3.11 575 4915 76 98.5 1290 0.13 0.16<br />

14 × 12 in 35.04 3.92 510 6763 79 98.5 1340 0.13 0.17<br />

16 × 14 in 39.37 2.81 450 8877 81 98.5 1357 0.133 0.15<br />

18 × 16 in 45.28 2.8 390 10859 83 98.5 1300 0.133 0.16<br />

20 × 18 in 55.12 2.76 340 13870 84 98.5 1281 0.119 0.16<br />

From Abulnaga (2001). Courtesy of Mazdak International Inc.<br />

TABLE 8-8 Recommendations for Im<strong>pe</strong>ller Diameter, S<strong>pe</strong>ed, S<strong>pe</strong>cific S<strong>pe</strong>ed Number,<br />

and Cavitations Parameter Metal-Lined or Hard Metal Mill Discharge Pumps (U.S.<br />

Units)<br />

Head Sigma<br />

Vane d2, d2/dS, S<strong>pe</strong>ed, Flow, Efficiency, Head, S<strong>pe</strong>cific factor, cavitation<br />

Model inch tip/suction rpm US gpm % ft s<strong>pe</strong>ed �US factor<br />

8 × 6 in 20.87 3.48 1005 2790 70 180 1186 0.173 0.14<br />

10 × 8 in 26.77 3.35 903 4721 74 180 1261 0.13 0.16<br />

12 × 10 in 31.10 3.11 779 6654 76 180 1290 0.13 0.16<br />

14 × 12 in 35.04 3.92 691 9121 79 180 1340 0.102 0.17<br />

16 × 14 in 39.37 2.81 609 12018 81 180 1357 0.132 0.15<br />

18 × 16 in 45.28 2.8 528 14701 83 180 1300 0.133 0.16<br />

20 × 18 in 55.12 2.76 460 18779 84 180 1281 0.12 0.16<br />

From Abulnaga (2001). Courtesy of Mazdak International Inc.<br />

The pump casing of a <strong>slurry</strong> pump often takes the sha<strong>pe</strong> of a volute. The best hydraulic<br />

design calls for a constant momentum design or a linear increase of the cross-sectional<br />

area from the tongue to the throat (Figure 8-19). In reality, the profile of the volute is often<br />

simplified to two semicircles. The idea is that hard metals are difficult to cast, and if<br />

the sha<strong>pe</strong> can be simplified, the casting will flow better during solidification.<br />

Rc in Figure 8-19 refers to the cutwater radius. The difference between Rc and R2 is effectively<br />

the gap at the cutwater. It must be large enough to accommodate the passage of<br />

coarse particles or rocks.<br />

The head develo<strong>pe</strong>d by the pump at shut-off is the sum of the head due to the rotation<br />

of the im<strong>pe</strong>ller and sha<strong>pe</strong> of the volute. Turton (1994) summarized the research of Frost<br />

and Nilsen (1991), who concluded that the shut-off head was insensitive to the number of<br />

blades, the blade outlet geometry, and the channel width of the im<strong>pe</strong>ller. They determined<br />

that:<br />

HSV = HIMP SV + HVOL SV<br />

(8-22)


8.26 CHAPTER EIGHT<br />

FIGURE 8-19 Parameters for the calculations of the shut-off head of a water pump used in<br />

Equation 8-24. (From Frost and Nilsen, 1991. Reproduced by <strong>pe</strong>rmission from the Institution<br />

of Mechanical Engineers, UK.)<br />

where<br />

HIMP SV = shut-off head due to the im<strong>pe</strong>ller<br />

HVOL SV = shut-off head due to the volute<br />

HSV = total shut-off head<br />

HIMP SV = [1 – (Rs/R2) 2 2 2 R2� � ] (8-23)<br />

2g<br />

and<br />

2<br />

HVOL SV = � ��R2 2 2<br />

R2� R4 – R2<br />

MD ln(R4/R2) – 2RMD(R4 – R2) + �<br />

� RMD – R 2<br />

� 2<br />

/g (8-24)<br />

Equations (8-23) and (8-24) were derived for water pumps, and it is recommended to<br />

confirm the results when designing a new family of <strong>slurry</strong> pumps.<br />

Referring to Figure 8-20, the width of the volute is defined by two components, X v in<br />

the x-direction and Y v in the y direction, when the volute is in a position for vertical top<br />

discharge. The magnitude of these two components de<strong>pe</strong>nds on the clearance at the cutwater,<br />

the throat area, the tip diameter of the im<strong>pe</strong>ller, and the discharge diameter of the<br />

pump. These are refined through ex<strong>pe</strong>rimental testing and hydraulic analysis. A good<br />

starting point (or rule of thumb) for the design engineer is to use the shroud diameter of<br />

the im<strong>pe</strong>ller d t as a reference and to establish<br />

X V = K xd t 1.3 < K x < 1.4 (8-25)<br />

Y V = K yd t 1.2 < K y < 1.3 (8-26)


cutwater<br />

t L<br />

(liner thickness)<br />

R 4<br />

THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

R 3<br />

R 2<br />

R 3<br />

X V<br />

R<br />

c<br />

R M<br />

R<br />

4<br />

discharge<br />

8.27<br />

im<strong>pe</strong>ller<br />

tip diameter<br />

throat area<br />

Having established a profile of the volute, the thickness of the liner and the thickness<br />

of the casing are then added before locating the bolts for lined casings.<br />

There is no definite rule of thumb for the thickness of rubber or metal liners. The<br />

thickness of the liner is established by the manufacturer on the basis on their ex<strong>pe</strong>rience<br />

with the application. Table 8-9 recommends a liner thickness in the range of 4% to 6% of<br />

the im<strong>pe</strong>ller diameter.<br />

Having sized the thickness of the liner, a parameter D for the volute is defined using<br />

the width XV as<br />

D = XV + 2tL (8-26)<br />

For a single-stage pump designed for a pressure of 1035 kPa (150 psi), with a ribbed<br />

casing, the casing thickness is established as<br />

tc � D/41 (8-27)<br />

Equation 8-27 should be complemented by a full finite element analysis, as the ribs<br />

have to be placed correctly. Modern computers are very useful for checking on the size of<br />

the ribs. Burgess and Abulnaga (1991) have recommended the use of the equivalent thickness<br />

approach. It consists of calculating the second moment of area of the ribs and implementing<br />

them in a plate model for the casing. An alternative but much more tedious approach<br />

is to use brick elements. Since 1991, the science of minicomputers has advanced<br />

greatly and it is now possible to implement very sophisticated three-dimensional models.<br />

t c<br />

suction diameter<br />

Y v<br />

(casing thickness)<br />

FIGURE 8-20 Volute sha<strong>pe</strong> of a <strong>slurry</strong> pump simplified for the sake of manufacturing and<br />

casting of hard metal casing or liners to a minimum number of partial circles.


8.28 CHAPTER EIGHT<br />

TABLE 8-9 Recommended Dimensions for a Single Stage Mill Discharge Pump (metric size<br />

example)<br />

Size (mm) 200 × 150 250 × 200 300 × 250 350 × 300 400 × 350 450 × 400 500 × 450<br />

Im<strong>pe</strong>ller d2 530 680 790 890 1000 1150 1400<br />

Shroud diameter dt 560 720 830 930 1050 1200 1500<br />

Cutwater diameter dC 657 843 980 1104 1240 1426 1775<br />

Cutwater gap<br />

(dC – dt)/2 49 62 75 87 95 113 138<br />

XV = 1.3 dt 728 936 1073 1209 1352 1560 1950<br />

YV = 1.25 dt 700 900 1031 1163 1300 1500 1875<br />

Liner thickness tL 34 38 41 45 48 51 55<br />

D = XV + 2 · tL 796 1012 1155 1299 1448 1662 2060<br />

Pressure area Ap (m2 )* 0.503 0.82 1.064 1.363 1.70 2.24 3.87<br />

Working pressure kPa 1035 1035 1035 1035 1035 1035 1035<br />

Design pressure kPa 1380 1380 1380 1380 1380 1380 1380<br />

F = Ap · Pdesign (kN) 694 1132 1468 1881 2348 3105 5341<br />

D/t 40 40.7 41.17 40.42 41 41.07 41.2<br />

Casing thickness<br />

tc (with ribs)<br />

20 24 28 31 34 39 50<br />

Number of bolts 12 12 12 12 12 12 12<br />

Load/bolt kN 58 94 122 157 196 259 445<br />

Bolt area mm2 ** 347 563 731 940 1174 1551 2662<br />

Bolt diameter mm 21 27 31 35 39 45 58<br />

Bolt M24 M30 M36 M40 M46 M50 M62<br />

*A p = 0.9[X V + t L][Y V + t L] · 10 –6 .<br />

**Allowed stress on bolt 166 Mpa.<br />

Note: these calculations are preliminary and must be confirmed by finite element analysis. They are for<br />

single stage, 1.38 MPa rating with ductile iron casing.<br />

Having established the thickness of the casing, it is important to establish the size and<br />

number of bolts for radial split casings. An equivalent pressure area is then established using<br />

the following formula:<br />

Ap = 0.9[XV + tL][YV + tL] (8-28)<br />

The design pressure PD is usually established as the maximum o<strong>pe</strong>rating pressure times a<br />

factor of 1.25. It is then multiplied by Ap to obtain the total force on the casing Fp: Fp = PD · Ap (8-29)<br />

The size and number of bolts is then established using the yield stress of the bolts.<br />

Detailed finite element analysis of multistage tailings pumps has demonstrated that the<br />

maximum stress occurs at the cutwater. Some of the very high pressure pumps feature a<br />

s<strong>pe</strong>cial bolt at the cutwater that is larger than the other bolts (Burgess and Abulnaga,<br />

1991).<br />

Table 8-9 presents some recommendation for average dimensions of a single-stage<br />

mill discharge pump designed for a maximum o<strong>pe</strong>rating pressure of 1035 kPa (150 psi).<br />

In this example, it was arbitrarily assumed that the number of bolts is 12, to give the reader<br />

an idea of the effect of loads on size of bolts. Obviously, on the larger pumps, the de-


THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

TABLE 8-10 Recommended Dimensions for a Single Stage Mill Discharge Pump<br />

(USCS units size)<br />

8.29<br />

Size (in) 8 × 6 10 × 8 12 × 10 14 × 12 16 × 14 18 × 16 20 × 18<br />

Im<strong>pe</strong>ller d2 21� 26.8� 31� 35� 39.4� 45.3� 55�<br />

Shroud diameter dt 22� 28.3� 32.7� 36.6� 41.3� 47.25� 59�<br />

Cutwater diameter dC 25.9� 33.2� 38.6� 43.5� 48.8� 56.1� 69.9�<br />

Cutwater gap (dC – dt)/2 1.95� 2.45� 2.95� 3.45� 3.77� 4.43� 5.45�<br />

XV = 1.3 dt 28.6� 36.8� 42.5� 47.6� 53.7� 61.43� 76.7�<br />

YV = 1.25 dt 27.5� 35.4� 40.9� 45.8� 51.6� 59� 73.8�<br />

Liner thickness tL 1.34� 1.5� 1.6� 1.77� 1.89� 2� 2.16�<br />

D = XV + 2 · tL 31.3� 39.8� 45.7� 51.1� 57.5� 65.43� 81�<br />

Pressure area Ap (in2 )* 777 1272 1687 2114 2973 3482 5990<br />

Working pressure psi 150 150 150 150 150 150 150<br />

Design pressure psi 200 200 200 200 200 200 200<br />

F = Ap*Pdesign (lbf) 155400 254389 337400 422800 594600 696400 1198000<br />

D/t 40 40.7 41.17 40.42 41 41.07 41.2<br />

Casing thickness tc (with ribs)<br />

0.78� 0.95� 1.1� 1.22� 1.34� 1.54� 2�<br />

Number of bolts 12 12 12 12 12 12 12<br />

Load/bolt lbf 12,950 21,199 28,117 35,233 49,550 58,033 99,833<br />

Bolt area in2 ** 0.539 0.883� 1.17 1.47 2.06 2.42 4.16<br />

Min Bolt diameter 0.83� 1.06 1.22� 1.38� 1.62� 1.75� 2.3�<br />

Bolt size (in) 7/8� 11/4 1.375 1.5� 1.75� 2� 2.5�<br />

*A p = 0.9[X V + t L][Y V + t L].<br />

**Allowed stress on bolt 24,000 psi.<br />

Note: these calculations are preliminary and must be confirmed by finite element analysis. They<br />

are for single stage, 200 psi rating with ductile iron casing.<br />

signer may increase the number of bolts to keep them within a reasonable size. Table 8-10<br />

is a similar table using USCS units.<br />

The casing pump takes the sha<strong>pe</strong> of the volute (Figure 8-21). In addition to the volute<br />

liner, a front wear plate or throatbush (Figure 8-22) is bolted to the casing.<br />

Compared to a water pump, a <strong>slurry</strong> pump has a much wider gap at the cutwater with res<strong>pe</strong>ct<br />

to the im<strong>pe</strong>ller. This is due to the fact that the <strong>slurry</strong> pump must move solids that<br />

should not jam at the cutwater. In certain cases, oversized pumps were sold to mines and recirculation<br />

problems develo<strong>pe</strong>d with excessive wear. Manufacturers have gone back over<br />

their designs and extended the cutwater to cut down the flow by creating a sort of throttling<br />

effect. They call this sort of volute a low- flow volute (Figure 8-23). The advantage of this<br />

approach is that the pattern of the liner can be modified without having to replace the casing<br />

of the pump. Installing a so-called “reduced eye” im<strong>pe</strong>ller may also complement this<br />

approach. A “reduced eye” im<strong>pe</strong>ller is an im<strong>pe</strong>ller with a suction diameter smaller than the<br />

suction diameter of the casing. This provides a way to throttle the suction. The throatbush<br />

of the pump must also be modified to accommodate the reduced eye of the im<strong>pe</strong>ller.<br />

In the case of water pumps, the emphasis is to o<strong>pe</strong>rate as close as possible to the best<br />

efficiency point, where losses are at a minimum. In the case of <strong>slurry</strong> pumps, the situation<br />

is more complex, as the best efficiency point does not necessarily coincide with the minimum<br />

wear point. Certain designs of <strong>slurry</strong> pumps do point to minimum wear at 80% of


8.30 CHAPTER EIGHT<br />

FIGURE 8-21 Casting for the casing and cover plate of a vertical sump pump—clearly<br />

showing the volute sha<strong>pe</strong>—with an integral cast elbow at the discharge. (Courtesy of Mazdak<br />

International Inc.)<br />

the best efficiency point. This point is too often overlooked when sizing pumps. The consultant<br />

engineer is encouraged to discuss this point with the manufacturer. Certain manufacturers<br />

of pumps have in-house computational fluid dynamics programs to do a wear<br />

<strong>pe</strong>rformance analysis. Unfortunately, too often these give a two-dimensional profile of<br />

velocity in the volute, but insufficient data about vortices in the corners where gouging<br />

wear may develop.<br />

FIGURE 8-22 Throatbush or suction liner fixed to the pump front casing plate of a horizontal<br />

pump. The casing sha<strong>pe</strong> indicates the volute sha<strong>pe</strong> of the liner.


solid<br />

passageway<br />

THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

original cutwater<br />

extended cutwater<br />

for "low flow" volute<br />

throat<br />

8.31<br />

modified throatbush<br />

reduced<br />

eye im<strong>pe</strong>ller<br />

FIGURE 8-23 Restraining the flow by extending the cutwater and modifying the throat of<br />

the volute or liner, or decreasing the suction diameter of the im<strong>pe</strong>ller are methods for correcting<br />

oversized pumps.<br />

Example 8-4<br />

A new mine requires a very large pump to handle 1514 L/s (24,000 US gpm), at a total<br />

dynamic head of 43 m (141 ft) and a s<strong>pe</strong>cific gravity of the mixture of 1.5. Establish some<br />

preliminary parameters of design for the casing prior to conducting a finite element analysis.<br />

The head ratio is assumed to be 0.9 (see Chapter 9). Assume that this is a pump designed<br />

for single-stage o<strong>pe</strong>ration with a design pressure of 1.4 MPa (200 psi).<br />

Solution in SI Units<br />

The equivalent water head is 43 m/0.9 = 47.8 m. This is therefore an application for an<br />

all-metal pump. Table 8-5 suggests an average suction s<strong>pe</strong>ed of 6 m/s and a discharge<br />

s<strong>pe</strong>ed of 9 m/s at a discharge head of 55 m. Since the pump will o<strong>pe</strong>rate at 47.8 m, the<br />

ratio of tip s<strong>pe</strong>eds is �(4�7�.8�/5�5�)� = �0�.8�6�8� = 0.932. The pump will o<strong>pe</strong>rate at 0.932 of<br />

the maximum allowed s<strong>pe</strong>ed of 38 m/s for all metal im<strong>pe</strong>llers, or 35.42 m/s (or 116<br />

ft/s):<br />

�SI = = = 0.187<br />

The flow suction s<strong>pe</strong>ed is established as the ratio of tip s<strong>pe</strong>ed. This ratio is 0.932, and<br />

using the suggested maximum s<strong>pe</strong>ed of 6 m/s for metal im<strong>pe</strong>llers, the suction s<strong>pe</strong>ed Vs at<br />

the flow rate of 1514 L/s is then 0.932 × 6 = 5.59 m/s.<br />

The suction area = Q/Vs = 1.514/5.59 = 0.271 m2 .<br />

The corresponding inner diameter is 0.587 m or 23.12�.<br />

The discharge s<strong>pe</strong>ed Vd is 0.932 × 9 = 8.4 m/s.<br />

The discharge area = Q/Vd = 1.514/8.4 = 0.18 m2 gHBEP 9.81 · 47.8<br />

� ��<br />

2 2<br />

2U 2 2 · 35.42<br />

.<br />

The discharge inner diameter is 0.478 m or 18.8�.<br />

These values of suction and discharge diameter will be added to the liner thickness<br />

and to the casing thickness before calculating suction and discharge flanges and their corresponding<br />

bolt circles.<br />

Using Table 8-8 as a reference, the tip-to-suction diameter of the im<strong>pe</strong>ller ratio is assumed<br />

to be 2.75, or the tip diameter of the im<strong>pe</strong>ller becomes 0.587 × 2.75 = 1.615 m.


8.32 CHAPTER EIGHT<br />

Since U = 35.42 m/s,<br />

� = U/R = 35.42/1.615 = 21.93 rad/s<br />

N = 21.93 · 60/(2 · �) = 209.4 rpm<br />

Let us round it to 210 rpm. From Equation 8-16, the s<strong>pe</strong>cific s<strong>pe</strong>ed (in the International<br />

System of Units) is<br />

N · �Q� 210 · �1�.5�1�4�<br />

Nq = � = �� = 14.22<br />

3/4 H<br />

47.8 3/4<br />

Table 8-9 recommends that the shroud diameter dt be about 6% larger than the im<strong>pe</strong>ller<br />

vane diameter dV or 1.06 · 1.615 = 1.712 m.<br />

The next step is to establish a preliminary layout of the volute using Equations 8-25<br />

and 8-26. It is assumed that<br />

Kx = 1.35<br />

or<br />

XV = 1.35 · 1.71 = 2.31 m<br />

and<br />

Ky = 1.25<br />

or<br />

XV = 1.25 · 1.71 = 2.14 m<br />

Table 8-9 recommends a liner thickness in the range of 4% to 6% of the im<strong>pe</strong>ller<br />

shroud diameter:<br />

tL = 0.04 · 1.712 = 0.0685 m<br />

Let us assume 69 mm. Having sized the thickness of the liner, a parameter D defined in<br />

Equation 8-26 is:<br />

D = 2.31 + 2 · 0.069 = 2.45 m<br />

For a well-ribbed casing, the thickness of the casing tC in a preliminary stage of analysis<br />

is D/40 or 2450/40 = 63.5 mm; let us assume 64 mm. The outer diameter of the suction<br />

nozzle is therefore<br />

587 mm + 2 · (69 + 64) = 853 mm or 33.5�<br />

This suggests further iteration or the installation of a companion flange to 900 mm for Euro<strong>pe</strong>an<br />

sizes of pi<strong>pe</strong>s or 36� suction pi<strong>pe</strong>s for U.S. sizes of pi<strong>pe</strong>s.<br />

The outer diameter of the discharge nozzle is therefore<br />

478 mm + 2 · (69 + 64) = 744 mm or 29.29�<br />

These calculations suggest that the pump is effectively a pump with a discharge flange<br />

of 750 mm for metric pi<strong>pe</strong> sizes or 30� for U.S. sizes of pumps. The equivalent pressure<br />

area Ap is then established using Equation 8-28:<br />

Ap = 0.9[XV + tL][YV + tL] = 0.9[2.31 + 0.069][2.14 + 0.069] = 4.13 m2 At a design pressure of 1.4 MPa, the total force that the bolts must retain is:<br />

Fp = Ap · 1.4 MPa = 4.13 · 1.4 = 5.78 MN


THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

Since this is a fairly large casing, the design engineer decides to try 24 bolts around the<br />

casing. Each bolt will retain 5.78 MN/24 = 0.241 MN or 241 kN, assuming an allowed<br />

stress on bolt of the order of 166 Mpa. The cross-sectional area of the bolt at the minimum<br />

thread diameter is 0.241/166 = 0.00145 m2 or a diameter of 42 mm. 20 M48 bolts<br />

are therefore recommended.<br />

Solution in USCS Units<br />

The equivalent water head is 141 ft/0.9 = 156.8 ft of water. This is therefore an application<br />

for an all-metal pump. Table 8-5 suggests an average suction s<strong>pe</strong>ed of 19.7 ft/s and a<br />

discharge s<strong>pe</strong>ed of 29.5 ft/s at a discharge head of 180.5 ft. Since the pump will o<strong>pe</strong>rate at<br />

47.8 m, the ratio of tip s<strong>pe</strong>eds is �(1�5�6�.8�/1�8�0�.5�) = �0�.8�6�8� = 0.932. The pump will o<strong>pe</strong>rate<br />

at 0.932 of the maximum allowed s<strong>pe</strong>ed of 124.67 ft/sec for all metal im<strong>pe</strong>llers, or 116<br />

ft/s:<br />

gHBEP 32.2 · 156.8<br />

�US = � = �� = 0.375<br />

2 2<br />

U 2 116<br />

The flow suction s<strong>pe</strong>ed is established as the ratio of tip s<strong>pe</strong>ed. This ratio is 0.932, and<br />

using the suggested maximum s<strong>pe</strong>ed of 19.7 ft/s for metal im<strong>pe</strong>llers, the suction s<strong>pe</strong>ed Vs at the flow rate of 24,000 US gpm (53.47 ft3 /sec) is then 0.932 × 19.7 = 18.36 ft/s.<br />

The suction area = Q/Vs = 53.47 ft3 /18.36 = 2.912 ft2 .<br />

The corresponding inner diameter is 1.926 ft or 23.12�.<br />

The discharge s<strong>pe</strong>ed Vd is 0.932 × 29.5 ft/s = 27.5 ft/sec.<br />

The discharge area = Q/Vd = 53.47/27.5 = 1.944 ft/sec2 .<br />

The discharge inner diameter is 1.573 ft or 18.9�.<br />

These values of suction and discharge diameter will be added to the liner thickness<br />

and to the casing thickness before calculating suction and discharge flanges and their corresponding<br />

bolt circles.<br />

Using Table 8-8 as a reference, the tip-to-suction diameter of the im<strong>pe</strong>ller ratio is assumed<br />

to be 2.75, or the tip diameter of the im<strong>pe</strong>ller becomes 23.12� × 2.75 = 63.6 in or<br />

5.3 ft.<br />

Since U = 116 ft/s,<br />

� = U/R = 116/5.3 = 21.9 rad/s<br />

N = 21.9 · 60/(2 · �) = 209.4 rpm<br />

Let us round it to 210 rpm. From Equation 8-16, the s<strong>pe</strong>cific s<strong>pe</strong>ed (In the International<br />

System of Units) is<br />

N · �Q� 210 · �2�4�0�0�0�<br />

NUS = � = �� = 734<br />

3/4 H<br />

156.8 3/4<br />

8.33<br />

Table 8-9 recommends that the shroud diameter dt be about 6% larger than the im<strong>pe</strong>ller<br />

vane diameter dV or 1.06 · 63.6� = 67.42�.<br />

The next step is to establish a preliminary layout of the volute using Equations 8-25<br />

and 8-26. It is assumed that<br />

Kx = 1.35<br />

or<br />

Xv = 1.35 · 67.42� = 91�<br />

and<br />

Ky = 1.25


8.34 CHAPTER EIGHT<br />

or<br />

Xv = 1.25 · 67.42� = 84.3�<br />

Table 8-9 recommends a liner thickness in the range of 4% to 6% of the im<strong>pe</strong>ller<br />

shroud diameter:<br />

tL = 0.04 · 67.42� = 2.69�<br />

Let us assume 2.7�. Having sized the thickness of the liner, a parameter D defined in<br />

Equation 8-26:<br />

D = 91� + 2 · 2.7� = 96.4�<br />

For a well-ribbed casing, the thickness of the casing tC in a preliminary stage of analysis<br />

is D/40 or 96.4�/40 = 2.41�. The outer diameter of the suction nozzle is therefore<br />

23.12� + 2 · (2.7� + 2.41�) = 33.34�<br />

This suggests further iteration or the installation of a companion flange to 36� suction<br />

pi<strong>pe</strong>s for U.S. sizes.<br />

The outer diameter of the discharge nozzle is therefore<br />

18.9� + 2 · (2.7� + 2.41�) = 29.12�<br />

These calculations suggest that the pump is effectively a pump with a discharge flange<br />

of 30� for U.S. sizes of pi<strong>pe</strong>s. The equivalent pressure area Ap is then established using<br />

Equation 8-28:<br />

Ap = 0.9[XV + tL][YV + tL] = 0.9[91 + 2.7][84.4 + 2.7] = 7345.14 in2 At a design pressure of 200 psi, the total force that the bolts must retain is:<br />

Fp = Ap · 1.4 MPa = 7345.14 · 200 = 1,469,028 lbf<br />

Since this is a fairly large casing, the design engineer decides to try 24 bolts around the<br />

casing. Each bolt will retain 1,469,028 lbf/24 = 61,210 lbf, assuming an allowed stress on<br />

bolt of the order of 24,000 psi. The cross-sectional area of the bolt at the minimum thread<br />

diameter is 61,210 lbf/24,000 = 2.55 in2 or a diameter of 1.8�. 20 1.875� bolts are therefore<br />

recommended.<br />

The design engineer must make allowance for the diameter of washers and the spotfacing<br />

diameter while laying down the design of the casing, as explained in Table 8-11.<br />

To complete this preliminary design exercise, the engineer needs to calculate the width of<br />

the im<strong>pe</strong>ller, including the pump-out vanes. This will be the topic of Section 8-4.<br />

8-4 THE IMPELLER, EXPELLER AND<br />

DYNAMIC SEAL<br />

Slurry, like any liquid, tends to find its way of least resistance. When a pressure difference<br />

exists between the volute pressure and the suction pressure at the front of a <strong>slurry</strong><br />

pump or the gland and stuffing box pressure (leaking to atmosphere) exits, <strong>slurry</strong> tends to<br />

flow back. However, as passageways narrow near the stuffing box or near the suction,<br />

solids become entrap<strong>pe</strong>d and accelerate abrasive wear.<br />

Leakage of <strong>slurry</strong> at the stuffing box can be dangerous to the environment, and can<br />

damage bearings. Various methods have been develo<strong>pe</strong>d over the years to counteract<br />

leaks. One popular method consists of injecting water at the gland. The gland water pres-


THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

TABLE 8-11 Size of Metric bolts and Allowance for Spot Facing. Suitable for Slurry<br />

Pump Casing and Stuffing Box<br />

sure is usually 35–70 kPa (5–10 psi) above the discharge pressure of the pump. The water<br />

acts also as a cooling lubricant to the shaft sleeve and packing rings. As time passes, the<br />

abradable packing rings wear slowly, and the o<strong>pe</strong>rator has to readjust the gland. Thus, the<br />

gland rings are usually split with tightening bolts (Figure 8-24).<br />

Unfortunately, trucking or pumping fresh gland water to remote tailing pump stations<br />

is not always the most economical solution. The pumping cost of gland water is not negligible<br />

for large pumps. In some cases such as pumping ore concentrate, the process engineer<br />

would prefer to avoid diluting the <strong>slurry</strong> by adding water at the gland. In the mid-<br />

1960s, <strong>slurry</strong> pump designers started to investigate the concept of a dynamic seal. A<br />

dynamic seal in its most basic concept consists of a ring of vanes on a shroud capable of<br />

creating a vortex. The designer of the dynamic seal tries to create a vortex field strong<br />

enough to prevent flow to the center of the vortex. In fact, when pressure is sufficiently<br />

reduced at the center of the dynamic seal to a magnitude below the outside atmospheric<br />

value, air is sucked in through the gland, and an air ring is formed .<br />

Despite the ap<strong>pe</strong>arance of ex<strong>pe</strong>llers, dynamic seals, and pump-out vanes in the mid-<br />

1960s, there is a dearth of technical information of their <strong>pe</strong>rformance. Various claims<br />

made in sales brochures are difficult to substantiate. Universities research centers have<br />

not paid much attention either. In some res<strong>pe</strong>cts, the ex<strong>pe</strong>ller at first look condradicts traditional<br />

thinking. It is in fact an im<strong>pe</strong>ller whose purpose is to re<strong>pe</strong>l or prevent flow. It<br />

goes against the logic of rotodynamics.<br />

The dynamic seal of a <strong>slurry</strong> pump consists of:<br />

� Pump-out vanes on the back shroud of the im<strong>pe</strong>ller (Figure 8-25)<br />

� Antiswirl vanes between the im<strong>pe</strong>ller and the ex<strong>pe</strong>ller<br />

� one or more ex<strong>pe</strong>llers with antiswirl vanes between them<br />

8.35<br />

Clearance hole Washer outside Spot facing Erix Back Spot<br />

Bolt size diameter (mm) diameter (mm) diameter (mm) facing diameter (mm)<br />

M5 6 10 12 15<br />

M6 7 12.5 14 15<br />

M8 9 17 19 18<br />

M10 12 21 24 24<br />

M16 18 30 33 33<br />

M20 23 37 41 43<br />

M24 27 44 46 48<br />

M30 33 56 60 62<br />

M36 39 66 70 72<br />

M42 45 78 80 82<br />

M48 51 92 96 108<br />

M56 59 105 110 113<br />

M64 67 115 120 122<br />

The dynamic seal o<strong>pe</strong>rates only when the pump is rotating at a sufficient s<strong>pe</strong>ed. When<br />

the pump is stationary, the dynamic seal ceases to <strong>pe</strong>rform and liquid may leak through<br />

the stuffing box, unless an additional stationary seal is provided or external water at sufficient<br />

pressure is flushing the gland.


8.36 CHAPTER EIGHT<br />

FIGURE 8-24 Stuffing box of the ZJ <strong>slurry</strong> pump (made in China) showing piping connection<br />

to inject water at high pressure and two adjusting bolts.<br />

FIGURE 8-25 Two front pump-out vanes of a <strong>slurry</strong> pump, before painting and testing (left)<br />

and painted with different colors (right), then installed in the pump of a test loop; the discoloration<br />

indicates patterns of wear. (Courtesy of Mazdak International, Inc.)


Flow in an ex<strong>pe</strong>ller is complex and de<strong>pe</strong>nds on the difference in relative motion between<br />

the stationary surface of the case liner and the rotating disk of the ex<strong>pe</strong>ller.<br />

Consider Figure 8-26 showing a closed im<strong>pe</strong>ller with pump-out vanes on the back<br />

shroud. The im<strong>pe</strong>ller main vane tip radius is R2, but the pump-out vanes extend only to<br />

the radius Rr. A shaft sleeve behind the im<strong>pe</strong>ller has Rs as a tip radius. In the front shroud<br />

of the im<strong>pe</strong>ller, another set of pump-out vanes extend to the radius Rf and provide dynamic<br />

sealing between the im<strong>pe</strong>ller and the throatbush to re<strong>pe</strong>l any solids that may tend to<br />

slip toward the suction (where the pressure is obviously lower). As the im<strong>pe</strong>ller rotates, a<br />

pressure field develops on the front shroud of the im<strong>pe</strong>ller due to the front pump-out<br />

vanes, and another pressure field develops behind the im<strong>pe</strong>ller due to the back pump-out<br />

vanes. In an ideal world, both fields should balance each other. In reality, wear of these<br />

vanes and the difference of clearance between the front and the back vanes with res<strong>pe</strong>ct to<br />

the casing or its liners tend to create an unbalance.<br />

In reference to Table 8-1, Case 7 for a forced vortex we have:<br />

� = C7 × R0 v0<br />

–1 V × Rv0 = C7<br />

P/� = C 7 2 · R 2 v0/(2 · g) + h 7<br />

Stepanoff (1993) stipulated that when a disk is rotating against a stationary surface,<br />

the average angular s<strong>pe</strong>ed of the liquid between the two is half the angular s<strong>pe</strong>ed of the<br />

disk. However, when vanes are added to the rotating disk, the rotational s<strong>pe</strong>ed of the liquid<br />

is expressed as<br />

Hvf<br />

R2<br />

THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

Rf<br />

R1<br />

8.37<br />

1 + t/x<br />

�liq = �imp��� (8-30)<br />

2<br />

x f<br />

t f<br />

B<br />

2<br />

FIGURE 8-26 Dynamic pressure distribution due to front and back pump-out vanes of a<br />

<strong>slurry</strong> pump im<strong>pe</strong>ller.<br />

t b<br />

xb<br />

R sl<br />

R r<br />

H vr


8.38 CHAPTER EIGHT<br />

where t is the depth of the pump-out vanes and x is the total gap between the im<strong>pe</strong>ller<br />

back shroud and the casing wear surface. x = s + t, where s is the gap between the pumpout<br />

vanes and the back shroud.<br />

Figure 8-27 represents a simplified case of pump-out vanes that extend down to the<br />

shaft sleeve diameter dsL. The average rotational s<strong>pe</strong>ed of the liquid between the rotating<br />

im<strong>pe</strong>ller and the stationary shroud therefore �imp/2. Applying the Euler head to this region,<br />

the head at the radius Rr is therefore:<br />

2 U n – Un–1<br />

�H = ��<br />

(8-31)<br />

2g<br />

2 2 �H = (R2 – R r) (8-32)<br />

Because vanes extend from Rr and Rsl, �<br />

2 2 �H = (R r – R sl) (8-33)<br />

So if H2 is the head at the tip of the im<strong>pe</strong>ller vane, then the head at the stuffing box (in the<br />

absence of any ex<strong>pe</strong>ller) is the head at the sleeve, or Hsl. Because a certain <strong>pe</strong>rcentage of<br />

the dynamic pressure is converted to static head in the volute, H2 is often assumed to be<br />

75% of the total dynamic head:<br />

2 imp(1 + t/x) 2<br />

�<br />

8g<br />

��<br />

8g<br />

� 2 imp<br />

� 2 imp<br />

Hsl = H2 – � � ([R2 2 � 2 2 2 2 – R r] – (1 + t/x) · (R r – R sl)) (8-34)<br />

8g<br />

The design engineer establishes H2 as a design criterion. Since the worst condition that<br />

a <strong>slurry</strong> pump may ex<strong>pe</strong>rience hap<strong>pe</strong>ns when it o<strong>pe</strong>rates at 30% of the B.E.P capacity and<br />

at a head H30, some engineers calculate H2 as:<br />

H2 = H30 – H1 When H s > H atm, the pump-out vanes will be completely flooded and the liquid will flow<br />

to the gland. To prevent this effect, some liquid at a higher pressure than the stuffing box<br />

pressure may be injected or an additional ex<strong>pe</strong>ller may be added. When H s < H atm, then<br />

the pump-out vanes suck in air and the stuffing box is sealed against loss of <strong>slurry</strong> (Figure<br />

8-26).<br />

In the back of the im<strong>pe</strong>ller, a second smaller disk with vanes facing the bearing assembly<br />

direction is sometimes installed (Figure 8-27). It is called the ex<strong>pe</strong>ller in the mining<br />

industry and the re<strong>pe</strong>ller in the pulp and pa<strong>pe</strong>r industry. Its diameter is usually smaller<br />

than 70% of the pump im<strong>pe</strong>ller. Its purpose is to reduce further the head between the hub<br />

of the im<strong>pe</strong>ller H b and the stuffing box.<br />

Equation (8-34) does not describe the effect of the number of vanes, the breadth of the<br />

vanes, or the sha<strong>pe</strong> of the vanes. Over the years, different manufacturers have develo<strong>pe</strong>d<br />

various sha<strong>pe</strong>s such as:<br />

� Straight radial vanes<br />

� Radial vanes but split in the middle with a gap<br />

� L-sha<strong>pe</strong>d vanes, also called hockey sticks<br />

� J-sha<strong>pe</strong>d vanes<br />

� Radial vanes with an outside ring


8.39<br />

c ve<br />

im<strong>pe</strong>ller<br />

Ød ho<br />

ex<strong>pe</strong>ller area<br />

lve<br />

LE<br />

t e<br />

h e<br />

Ød he<br />

FIGURE 8-27 Geometry of an ex<strong>pe</strong>ller with radial vanes.<br />

Exp<br />

Ød


8.40 CHAPTER EIGHT<br />

� Radial vanes with an outside ring and a middle ring<br />

� Lotus-sha<strong>pe</strong>d vanes<br />

These sha<strong>pe</strong>s are represented in Figure 8-28.<br />

Equation (8-34) clearly indicates that the head is proportional to the square of the<br />

s<strong>pe</strong>ed. There is therefore a minimum rotational s<strong>pe</strong>ed before that the dynamic seal starts<br />

to function.<br />

The consumed power of an ex<strong>pe</strong>ller is expressed as:<br />

P (kW) = constant · � · D5 · N3 (8-35)<br />

(a) backward curved vanes (b) radial split at midradius (c) L-sha<strong>pe</strong>d vanes ( hockey sticks)<br />

(e) simple radial<br />

(d) radial with ring at mid- radius<br />

(f ) lotus vanes<br />

FIGURE 8-28 Different sha<strong>pe</strong>s of vanes and rings of ex<strong>pe</strong>llers and dynamic seals.


THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

8.41<br />

Although various claims have been made in sales brochures about the merits of each<br />

vane ty<strong>pe</strong>, and numerous patents have been filed, there has been no substantial scientific<br />

data to confirm the claims. Often, the final sha<strong>pe</strong> is a compromise between the requirements<br />

for casting in hard metals and the requirements of the hydraulics.<br />

Im<strong>pe</strong>llers of <strong>slurry</strong> pumps must accommodate solids, and this means that the vanes must<br />

be wide enough. Each manufacturer has their own criteria, with dredge and gravel pumps<br />

requiring very wide im<strong>pe</strong>llers (Table 8-12). Adding this passageway to the thickness of the<br />

shrouds of pump-out vanes results in the im<strong>pe</strong>ller overall width b 2 (Figure 8-29).<br />

In Equation 8-35, it was pointed out that the power consumption from pump-out vanes<br />

is proportional to the diameter raised to the power of five. Instead of trimming the pumpout<br />

vanes to a diameter smaller than the im<strong>pe</strong>ller main vanes, they are sometimes ta<strong>pe</strong>red<br />

(t b and t f are gradually reduced toward the tip of the im<strong>pe</strong>ller; see Figure 8-29).<br />

In Figure 8-29, the pump-out vane thickness at the root is (g f + t fv), whereas at the tip it<br />

is t fv. In the back of the im<strong>pe</strong>ller, the pump-out vanes start at a diameter d b, whereas on the<br />

front side they start at d r. These values are plugged into Equation 8-34 to obtain R r in each<br />

case and to calculate axial thrust.<br />

Because <strong>slurry</strong> pumps are often cast in brittle alloys such as the high-chrome white<br />

iron, it is important to eliminate sharp edges that may act as stress risers. The manufacturers<br />

establish the radii R 3, R 4, R c, R r, R h, and R sv shown in Figure 8-29 to allow a smooth<br />

casting, but also to improve on the hydraulics. The effect of each parameter on the hydraulics<br />

as described in sales brochures is not always well proven.<br />

The vane diameter d 2 shown in Figure 8-29 is smaller than the shroud diameter d t, but<br />

it is the reference diameter for all calculations.<br />

The shaft sleeve with a diameter d sl is used in all thrust calculations. The sleeve protects<br />

the shaft from wear by the packing and solids that may accumulate between the<br />

packing rings.<br />

TABLE 8-12 Recommended Maximum Size of Spheres for the Design of the Width<br />

of Vanes of Slurry and Dredge Pumps<br />

Mill discharge pumps Gravel and dredge pumps<br />

_______________________________________ _______________________________________<br />

Discharge Size Sphere diameter, Discharge Size Sphere diameter,<br />

(mm) (inches) mm (in) (mm) (inches) mm (in)<br />

25 1.5 × 1 13 (1/2�)<br />

38 2 × 1 18 (11/16�)<br />

50 3 × 2 20 (3/4�)<br />

75 4 × 3 22 (7/8�)<br />

100 6 × 4 38 (�1.5�) 100 6 × 4 80 (�3)<br />

150 8 × 6 50 (�2�) 150 8 × 6 127 (�5�)<br />

200 10 × 8 63 (�2.5�) 200 10 × 8 180 (�7�)<br />

250 12 × 10 80 (�3) 250 12 × 10 230 (�9�)<br />

300 14 × 12 88 (�3.5�) 300 14 × 12 240 (�9.5�)<br />

350 16 × 14 100 (�4�) 350 16 × 14 250 (�10�)<br />

400 18 × 16 115 (�4.5�) 400 18 × 16 280 (�11�)<br />

450 20 × 18 127 (�5�) 450 20 × 18 305 (�12�)<br />

500 24 × 20 140 (�5.5�) 500 24 × 20 360 (�14�)<br />

600 28 × 24 150 (�6�) 600 28 × 24 380 (�15�)<br />

650 30 × 26 180 (�7�) 650 30 × 26 450 (�18�)<br />

915 40 × 36 530 (�21�)


8.42 CHAPTER EIGHT<br />

Ø d b<br />

Ø d th<br />

Ø d h<br />

Ø d sl<br />

t<br />

bs<br />

R<br />

h<br />

R sv<br />

t bv<br />

bv<br />

Most <strong>slurry</strong> pumps use a threaded shaft. The length of the shaft thread L th is used in<br />

calculations of axial load transmitted from the torque. Some pumps use BSW and others<br />

use ACME thread, and some manufacturers have also their own thread designs to make it<br />

difficult to pirate their im<strong>pe</strong>llers.<br />

It is important to establish the center of gravity of the im<strong>pe</strong>ller. In the absence of data,<br />

it is often assumed to be at a distance L h. It is also assumed in the calculations that the radial<br />

thrust force is applied at the same point.<br />

8-5 DESIGN OF THE DRIVE END<br />

g b<br />

b<br />

2<br />

The hydraulic loads from the pump wet end are ultimately transmitted to the pump shaft<br />

and bearings. Because of the need to access all the pump parts for replacement due to<br />

wear during maintenance, <strong>slurry</strong> pumps have standardized cantilever designs, with all<br />

bearings well protected from solids ingestion.<br />

The main loads that are transmitted to the pump shaft are:<br />

t<br />

fs<br />

t fv<br />

g<br />

f<br />

R<br />

R<br />

tb fv<br />

R R<br />

R<br />

2<br />

t<br />

c<br />

L th<br />

Ø d 1<br />

� Radial force due to pressure distribution in the volute<br />

� Axial force due to the pump-out vanes and ex<strong>pe</strong>llers<br />

h<br />

i<br />

R r<br />

fsv<br />

Ø d r<br />

FIGURE 8-29 Cross-section of an im<strong>pe</strong>ller for a <strong>slurry</strong> pump showing different geometrical<br />

parameters.


THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

� Weight of the im<strong>pe</strong>ller and ex<strong>pe</strong>ller<br />

� Torque due to s<strong>pe</strong>ed and power consumption<br />

� Radial force on the drive end from pulleys<br />

8-5-1 Radial Thrust Due to Total Dynamic Head<br />

The radial force is due to the uneven pressure distribution in the pump casing. It is expressed<br />

as:<br />

FR = K · �gHd2 · B2 (8-36)<br />

where<br />

d2 = tip diameter of the im<strong>pe</strong>ller vanes<br />

B2 = width of the pump casing<br />

As shown in Figure 8-26,<br />

B2 = b2 + xf + xb (8-37)<br />

Wear can chip at the surface of the im<strong>pe</strong>ller or the casing, thus causing an increase of<br />

xf and xb and a reduction of b2 through the life of the pump.<br />

The value of K may be as high as 0.40 near the shut-off head and as low as 0.10 at the<br />

best efficiency point. It is, however, recommended to conduct pro<strong>pe</strong>r measurements with<br />

proximity probes over the envelo<strong>pe</strong> of the flow rate during the design of a new pump. The<br />

proximity probes are used to measure the deflection at the gland. The magnitude of the<br />

force is then calculated from cantilever stress theory.<br />

As shown in Figure 8-30, different sha<strong>pe</strong>s of volutes give different values for the radial<br />

load. Stepanoff (1993) clearly indicated that the direction of the radial force reverses<br />

after the best efficiency point, whereas Angle et al. (1997) do not seem to agree with this<br />

supposition. A misunderstanding of the direction of this hydraulic radial force leads to totally<br />

different estimation of the bearing life. A calculation that assumes a zero radial load<br />

near the best efficiency point (following the Stepanoff approach) can lead to a bearing life<br />

ten times as high as another calculation that assumes that the same radial load adds to the<br />

weight of the im<strong>pe</strong>ller, creating a large bending moment on the shaft and reaction loads at<br />

the bearings. A smart salesman may try to convince the consultant <strong>slurry</strong> engineer of the<br />

su<strong>pe</strong>riority of his product over the com<strong>pe</strong>tition in terms of the rigidity of the bearing assembly,<br />

whereas in reality it is a matter of adding or subtracting loads.<br />

Shafts of <strong>slurry</strong> pumps have broken at the shaft thread, simply because the radial load<br />

was too high and caused rapid fatigue failure. It is therefore strongly recommended to<br />

limit the minimum flow rate to half the best efficiency flow rate at the given s<strong>pe</strong>ed. Throttling<br />

an oversize pump is not recommended at all. Downsizing or reducing the s<strong>pe</strong>ed of<br />

the pump is essential to avoid excessive radial load on the pump shaft.<br />

Each manufacturer has their recommended value of K for the calculation of the radial<br />

load and the bearing life.<br />

8-5-2 Axial Thrust Due to Pressure<br />

8.43<br />

The axial thrust is due to the fact that the pressure on the suction side is different from the<br />

pressure on the back of the im<strong>pe</strong>ller. There is a difference between plain im<strong>pe</strong>llers and<br />

im<strong>pe</strong>llers with pump-out vanes, but since pump-out vanes wear out with time due to abrasion<br />

and erosion, the design engineer should conduct his calculations for both cases of im-


8.44<br />

Head<br />

(a) true volute (b) two semi-circle casing (b) circular casing<br />

FR<br />

Q N<br />

Flow rate<br />

After<br />

Angle &<br />

Rudonov<br />

(1999)<br />

After<br />

Stepanoff<br />

(1993)<br />

Head<br />

FR<br />

Q<br />

N<br />

Flow rate<br />

After<br />

Angle &<br />

Rudonov<br />

(1999)<br />

After<br />

Stepanoff<br />

(1993)<br />

FIGURE 8-30 Radial load for different sha<strong>pe</strong>s of casing versus flow rate.<br />

Head<br />

F<br />

R<br />

Q<br />

N<br />

Flow rate


THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

<strong>pe</strong>llers with and without pump-out vanes. The presence of an ex<strong>pe</strong>ller or the addition of<br />

pressurized gland water does affect the axial thrust.<br />

Consider in Figure 8-31 a closed im<strong>pe</strong>ller without pump-out vanes. The pressure on<br />

the suction side is P s and at the suction diameter d 1. The pressure on the back of the im<strong>pe</strong>ller<br />

is P 1. The pressure above d 1 on both sides of the im<strong>pe</strong>ller is equal and balances out.<br />

In the back of the shaft sleeve and shaft there is atmospheric pressure P A, so the resultant<br />

force based on the shaft sleeve diameter is:<br />

T SL = 0.25�d 2 SLP A<br />

On the suction side, there is suction pressure P s, so the thrust force is:<br />

T S = 0.25�d 1 2 Ps<br />

8.45<br />

The net thrust is:<br />

2 2 FA = 0.25�{P1[d 1 – d SL] + PAd 2 2<br />

SL – Psd 1} (8-38)<br />

For the first stage, PS is calculated in a very similar way to the NPSH.<br />

Some manufacturers design the bearing assembly to absorb the axial thrust from a single<br />

stage and others standardize on three stages because they anticipate use in a wide<br />

range of applications from mill discharge to tailings disposal.<br />

Because tailings pumps are often used in series, the bearing assembly may be designed<br />

for a suction pressure equal to the discharge pressure of the stage before the last one, i.e.,<br />

if M is the number of stages:<br />

Ps = (M – 1)�g(TDHst) + PA (8-39)<br />

where TDHst is total dynamic head <strong>pe</strong>r stage.<br />

Referring to Figure 8-29, when pump-out vanes are added in the back shroud, Equation<br />

8-34 is then used to calculate the value of Pb at the root of the pump out vanes Rb: Pb = P2 – 0.125��2 2 2<br />

imp{[R2 – R b] – [(1 + tb/xb) 2 2 2 · (R2 – R b)]} (8-40)<br />

where P2 = 0.75�g(TDH) + PS. Ps<br />

R1<br />

d A d P A<br />

FIGURE 8-31 Axial loads on an im<strong>pe</strong>ller with plain shrouds.<br />

P 1<br />

sl


8.46 CHAPTER EIGHT<br />

The average thrust force on the back shroud of the im<strong>pe</strong>ller is<br />

2 2 T2b = 0.5(P2 + Pb) �{[R 2 – Rb] (8-41)<br />

This value of the pressure Pb is transmitted to the ex<strong>pe</strong>ller box and becomes the pressure<br />

at the ex<strong>pe</strong>ller tip diameter dexp (Figure 8-27). The pressure at the ex<strong>pe</strong>ller diameter dhe (which is often equal to the shaft sleeve or the pressure at the gland) is then<br />

Phe = Pb – 0.125��2 imp{[R2 exp – R2 he] – [(1 + te/(te + cve)) 2 · (R2 exp – R 2 he)]} (8-42)<br />

The average thrust force on the back shroud of the ex<strong>pe</strong>ller is<br />

Tbe = 0.5(Phe + Pb) �{R2 exp – R 2 he} (8-43)<br />

If the ex<strong>pe</strong>ller hub diameter is larger than the shaft sleeve, there is a component of axial<br />

thrust as<br />

Tesl = 0.5(Phe + PA) �{R2 he – R 2 SL} (8-44)<br />

On the back of the sleeve and shaft, the pressure is essentially atmospheric so that the<br />

thrust is<br />

Tsl = PA�R 2 SL<br />

(8-45)<br />

On the front shroud of the im<strong>pe</strong>ller, pump-out vanes are also added with some im<strong>pe</strong>llers.<br />

Applying Equation 8-34 to Figure 8-29, the pressure at the front hub Rr is therefore:<br />

Pr = P2 – 0.125��2 2 2<br />

imp{[R2 – R r] – [(1 + tf/xf) 2 2 2 · (R2 – R r)]} (8-46)<br />

The average thrust force on the front shroud of the im<strong>pe</strong>ller between R2 and Rr is:<br />

2 2 T2r = 0.5(P2 + Pr) �{[R2 – R r] (8-47)<br />

If the front shroud hub diameter dr is larger than the suction diameter ds, there is a component<br />

of axial thrust as<br />

2 2 Trs = 0.5 (Pr + Ps) �{R r – RS} (8-48)<br />

The thrust due to the suction pressure is then<br />

2 Ts = Ps�RS (8-49)<br />

Total axial thrust equals total thrust on the back shroud minus total thrust on the suction:<br />

FA = [t2b + Tbe + Tsl] – [Ts + Trs + T2r] (8-50)<br />

In multistage applications with a number of pump in series, the total axial thrust can<br />

change direction as the suction pressure is higher than atmospheric pressure, and the ex<strong>pe</strong>ller<br />

and pump-out vanes’ effectiveness in balancing thrust drops with increasing number<br />

of stages.<br />

Since the flow calculations need to be re<strong>pe</strong>ated at various points on the pump curve, a<br />

computer program would be useful. The program AXIAL-RADIAL was develo<strong>pe</strong>d by<br />

the author in Qbasic, a language easy to understand by most engineers, but ex<strong>pe</strong>rts may<br />

modify it to PASCAL, C+, Fortran, or other languages as it suits their needs. It calculates<br />

both hydraulic and axial loads on the pump im<strong>pe</strong>ller.<br />

COMPUTER PROGRAM “AXIAL-RADIAL”<br />

9 CLS<br />

REM calculations of axial and radial loads on a pump im<strong>pe</strong>ller<br />

pi = 4 * ATN(1)<br />

Rem Calculations will be done assuming a s<strong>pe</strong>cific gravity of 1.7


THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

sg = 1.7<br />

INPUT “model name “; na$<br />

INPUT “tip shroud diameter dt (mm) “; dt<br />

INPUT “vane tip diameter d2 (mm) “; d2<br />

INPUT “suction diameter ds (mm) “; ds<br />

INPUT “starting diameter for front pump out vanes dr (mm) “; dr<br />

INPUT “starting diameter for back pump out vanes db (mm) “; db<br />

INPUT “back hub diameter dh (mm) “; dh<br />

INPUT “shaft sleeve o.d dsl (mm) “; dsl<br />

INPUT “overall width of im<strong>pe</strong>ller B2 (mm) “; bx<br />

INPUT “ vane tip width b2 (mm) “; b2<br />

INPUT “thickness of front shroud tfs (mm)”; tfs<br />

INPUT “thickness of front pump out vanes tfv (mm) “; tfv<br />

INPUT “anticipated front gap (mm)”; gf<br />

sf = gf + tfv<br />

‘INPUT “thickness of back shroud tbs (mm)”; tbs<br />

INPUT “thickness of back pump out vanes tbv (mm) “; tbv<br />

INPUT “anticipated back gap (mm)”; gb<br />

sb = gb + tbv<br />

INPUT “s<strong>pe</strong>ed for metal version”; n<br />

PRINT “it shall be assumed that pump out vane to gap ratio =0.7”<br />

PRINT<br />

a1 = .25 * pi * (dr/25.4) ^ 2<br />

a2 = .25 * pi * (d2/25.4) ^ 2<br />

a3 = .25 * pi * (dsl/25.4) ^ 2<br />

a4 = .25 * pi * (ds/25.4) ^ 2<br />

a5 = .25 * pi * (db/25.4) ^ 2<br />

c = 25.4<br />

DIM h(10), fa(10), fan(10), nr(10), Q(10), k(10),fr(10),f(10)<br />

Rem assume a typical curve for an all metal im<strong>pe</strong>ller<br />

h(1) = 64;k(1)=0.4<br />

h(2) = 62.7;k(2)=0.35<br />

h(3) = 60.5;k(3)=0.25<br />

h(4) = 55;k(4)=0.15<br />

h(5) = 49.5;k(5)=0.10<br />

h(6) = 35;k(6)=0.12<br />

h(7) = 34.2;k(7)=0.15<br />

h(8) = 33;k(8)=0.20<br />

h(9) = 30;k(9)=0.22<br />

h(10) = 27,k(10)=0.25<br />

INPUT “best efficiency flow rate for metal version “; qnm<br />

Q(1) = .25 * qnm<br />

Q(2) = .5 * qnm<br />

Q(3) = .75 * qnm<br />

Q(4) = 1 * qnm<br />

Q(5) = 1.15 * qnm<br />

Rem calculation for rubber<br />

Q(6) = .25/1.354 * qnm<br />

Q(7) = .5/1.354 * qnm<br />

Q(8) = .75/1.354 * qnm<br />

Q(9) = 1/1.354 * qnm<br />

8.47


8.48 CHAPTER EIGHT<br />

Q(10) = 1.15/1.354 * qnm<br />

FOR i = 1 TO 10<br />

h = h(i)<br />

h2 = .8 * h/.3048<br />

PRINT “h2= “; h2<br />

INPUT “hit any key to continue “; l$<br />

IF h(i) > 35 THEN nr(i) = n<br />

IF h(i)


THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

TABLE 8-13 Limiting Dimensions of American National Standard General Purpose<br />

Single Start ACME Threads. External Threads (for Shafts), Class 2G<br />

8.49<br />

Nominal Threads Major diameter, Minor diameter, Pitch diameter,<br />

diameter (inch) <strong>pe</strong>r inch min/max, in inch min/max, in inch min/max, in inch<br />

1.50 4 1.4875–1.5000 1.1965–1.2300 1.3429–1.3652<br />

1.75 4 1.7375–1.7500 1.4456–1.4800 1.5916–1.6145<br />

2.00 4 1.9875–2.0000 1.6948–1.7300 1.8402–1.8637<br />

2.25 3 2.2333–2.2500 1.8572–1.8967 2.0450–2.0713<br />

2.50 3 2.4833–2.5000 2.1065–2.1467 2.2939–2.3207<br />

2.75 3 2.7333–2.7500 2.3558–2.3967 2.5427–2.5700<br />

3.00 2 2.9750–3.0000 2.4326–2.4800 2.7044–2.7360<br />

3.50 2 3.4750–3.5000 2.9314–2.9800 3.2026–3.2350<br />

4.00 2 3.9750–4.0000 3.4302–3.4800 3.7008–3.7340<br />

4.50 2 4.4750–4.5000 3.9291–3.9800 4.1991–4.2330<br />

5.00 2 4.9750–5.0000 4.4281–4.4800 4.6973–4.7319<br />

For more information consult ANSI standard B1.5-1977.<br />

ACME external thread (Table 8-13) is used for the shaft and the ACME internal thread<br />

(Table 8-14) is used for the im<strong>pe</strong>ller. Because the im<strong>pe</strong>ller thread is cast, particularly with<br />

hard metals, Class 2G is suggested because it has a wider range of tolerances than the 3G,<br />

4G, and 5G Classes. BSW shaft threads are used on the smallest sizes. Figure 8-32 represents<br />

a typical ACME shaft thread.<br />

In order to determine the shaft stresses and the axial pull due to torque, the first step is<br />

to assess the torque due power:<br />

Tq = 60PW/(2�N) (8-51)<br />

Example 8-5<br />

A pump is sized for 200 m 3 /hr, at a TDH of 36 m and s<strong>pe</strong>cific gravity of 1.4. The pump<br />

s<strong>pe</strong>ed is 600 rpm and the hydraulic efficiency is 67%. Determine the power and the torque.<br />

TABLE 8-14 Limiting Dimensions of American National Standard General Purpose<br />

Single Start ACME Threads, Internal Threads (for Im<strong>pe</strong>llers), Class 2G<br />

Nominal Threads Major diameter, Minor diameter, Pitch diameter,<br />

diameter (inch) <strong>pe</strong>r inch min/max, in inch min/max, in inch min/max, in inch<br />

1.50 4 1.5200–1.5400 1.2500–1.2625 1.3750–1.3973<br />

1.75 4 1.7700–1.7900 1.5000–1.5125 1.6250–1.6479<br />

2.00 4 2.0200–2.0400 1.7500–1.7625 1.8750–1.8985<br />

2.25 3 2.2700–2.2900 1.9167–1.9334 2.0833–2.1096<br />

2.50 3 2.5200–2.5400 2.1667–2.1834 2.3333–2.3601<br />

2.75 3 2.7700–2.7900 2.4167–2.4334 2.5833–2.6106<br />

3.00 2 3.0200–3.0400 2.5000–2.5250 2.7500–2.7816<br />

3.50 2 3.5200–3.5400 3.0000–3.0250 3.2500–3.2824<br />

4.00 2 4.0200–4.0400 3.5000–3.5250 3.7500–3.7832<br />

4.50 2 4.5200–4.5400 4.0000–4.0250 4.2500–4.2839<br />

5.00 2 5.0200–5.0400 4.5000–4.5250 4.7500–4.7846


8.50 CHAPTER EIGHT<br />

major dia<br />

pitch dia<br />

minor dia<br />

p/2<br />

b t<br />

pitch<br />

p<br />

2 = 29˚<br />

Solution in SI Units<br />

power = (200/3600) · 1.4 · 9810 · 36/0.67 = 40,997 W<br />

torque = power/rotational s<strong>pe</strong>ed = 40,997/(2 · � · N/60) = 652.5 N-m<br />

The helix angle � of the thread is defined as<br />

tan � = � � (8-52)<br />

where<br />

L = length of a full turn = pitch for single-start threads<br />

L = 2 × pitch for double start threads<br />

pitch = distance between two consecutive threads measured at the thread diameter<br />

dm = pitch diameter<br />

The axial load transmitted through the thread from the torque is expressed as<br />

�dm cos �n – fL<br />

Fth = 2 · Tq����� (8-53)<br />

dm( f�dm + L cos �n) tan �n = tan � cos �<br />

For ACME threads it equals 14.5°. For square threads it is nil. For modified square it is<br />

5°. For buttress threads it is 7°. So for an ACME thread:<br />

tan �n = 0.968 cos �<br />

The coefficient of friction f is measured between the shaft and the im<strong>pe</strong>ller. In some<br />

pumps, the shaft is of steel but the im<strong>pe</strong>ller may be of bronze. Slurry pumps are essentially<br />

steel against iron and the coefficient of friction is considered to be in the range of 0.14<br />

to 0.15:<br />

�dm cos �n – fL<br />

Fth = 2 · Tq����� (8-54)<br />

dm( f�dm + L cos �n) If n is the number of engaged threads, the axial load from this thread pull force creates a<br />

bending stress Sb and a shear stress Ss at the root of the shaft thread:<br />

L<br />

� �dm<br />

h = p/2<br />

FIGURE 8-32 ACME thread for pump shafts.


THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

3F thh<br />

� �dmnb t 2<br />

S b = � � (8-55)<br />

Ss = ��� (8-56)<br />

�dmnb t<br />

where<br />

h = height of the thread tooth = (major diameter – minor diameter)/2; in the case of<br />

ACME threads h = p/2<br />

bt = thread width at the root<br />

8-5-4 Radial Force on the Drive End<br />

When pulleys are installed to drive the pump, the torque transmitted through a pulley diameter<br />

D p results in a force. Different equations are available, but the simplest expresses<br />

the resultant pulley force as:<br />

4Tq �<br />

Dp<br />

8-5-5 Total Forces from the Wet End<br />

F p = � � (8-57)<br />

The total radial force transmitted by the im<strong>pe</strong>ller to the shaft is due to the combination of<br />

the hydraulic radial thrust and the weight of the im<strong>pe</strong>ller:<br />

F1 = FR + Wimp (8-58)<br />

It is assumed that F1 is acting on the center of gravity of the im<strong>pe</strong>ller. The total axial load<br />

is:<br />

F 2 = ±F A<br />

as the axial force may change direction as the number of stages exceeds two pumps in series.<br />

The torque, a source of torsion stress, was defined in Equation 8-51. On the drive<br />

side, the pulley force is upward for overhead-mounted motors or sideways for sidemounted<br />

motors. Calculations are often made on the assumption of overhead-mounted<br />

motors:<br />

F3 = Fp – Wp (8-59)<br />

On this basis, the shaft of the pump is designed. Due to fatigue considerations, the maximum<br />

stress should be smaller than the lesser of 18% yield, or 30% ultimate tensile<br />

strength.<br />

Referring to Figure 8-33, the equilibrium of forces shows that the reaction force at the<br />

wet end is RW and at the drive end it is RD: R W – F 1 – R D + F 3 = 0 (8-60)<br />

Taking moments at the point of contact load of the wet end bearing:<br />

–A · F1 + RD · B – F3(B + C) = 0<br />

F th<br />

8.51


8.52 CHAPTER EIGHT<br />

Torque<br />

F A<br />

L th<br />

F + W<br />

R imp<br />

F3(B + C)<br />

RD = � �<br />

F3C RW = � �<br />

A · F1 + � (8-61)<br />

B<br />

F1 ·(A + B)<br />

� + �� (8-62)<br />

B B<br />

The reader should refer to s<strong>pe</strong>cialized books on machine design that detail all as<strong>pe</strong>cts<br />

of the design of shafts, stress concentration, and bearing life calculations from the reaction<br />

forces at both the drive end (outboard) and wet end (inboard) bearings. The manufacturers<br />

of bearings have their own detailed factors for ty<strong>pe</strong> of lubricant and ratio of axial to<br />

radial force. Some manufacturers of <strong>slurry</strong> pumps offer grease lubricated bearing assemblies<br />

and reserve the oil version for high-s<strong>pe</strong>ed and high-thrust loads (as in pumps in series),<br />

whereas some use oil all across their range of pumps.<br />

8-5-6 Flange Loads<br />

d A<br />

d<br />

R<br />

W<br />

�� B<br />

WE<br />

d<br />

F p - Wpulley<br />

(with belt drive)<br />

A common misconception is that the flanges of <strong>slurry</strong> pumps can take the same loads as<br />

water pumps. The fact that the discharge flange is split radially to allow access to rubber<br />

or metal liners by itself is an indication that this is not the case at all. The casing of a <strong>slurry</strong><br />

pump can be distorted by excessive pi<strong>pe</strong> loads on the flange. The consultant engineer is<br />

therefore well advised to contact the manufacturer for allowed flange loads. It is also necessary<br />

to provide pro<strong>pe</strong>r pi<strong>pe</strong> supports at the discharge of the <strong>slurry</strong> pump, and not to use<br />

the pump by itself as an anchor block to piping.<br />

The common error is to apply a large expansion at the discharge of the pump, such as<br />

from a 4� pump discharge to an 8� pi<strong>pe</strong>. Doubling the diameter is effectively multiplying<br />

by four the area exposed to the full pressure, of which a quarter is absorbed by the pump,<br />

leaving three quarters to be balanced by a pi<strong>pe</strong> fitting such as a pro<strong>pe</strong>rly supported dead<br />

end bulkhead, an anchor block, or, whenever possible, by soil friction, as is the case with<br />

pi<strong>pe</strong>line pumps.<br />

R<br />

D<br />

A B C<br />

FIGURE 8-33 Loads on the shaft of a horizontal <strong>slurry</strong> pump.<br />

d k d D.E


THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

8-6 ADJUSTMENT OF THE WET END<br />

The wear of the im<strong>pe</strong>ller front vanes, the throatbush, is believed to cause a drop in efficiency.<br />

The im<strong>pe</strong>ller must be readjusted by moving it forward. To move the im<strong>pe</strong>ller relative<br />

to the casing, the shaft assembly must be moved relative to the pump frame, as the<br />

latter is bolted to the pump casing. Two different methods are available:<br />

1. A s<strong>pe</strong>cial bolt under the bearing cartridge (Figure 8-34)<br />

2. Push and pull bolts at the drive end (Figure 8-35)<br />

Once the bearing cartridge is moved, it is fixed in place by clamping bolts that are tightened<br />

against the frame.<br />

8-7 VERTICAL SLURRY PUMPS<br />

8.53<br />

The vertical sump (Figure 8-36) complements the horizontal pump. The vertical pump is<br />

particularly suitable for floor sumps in mill discharge areas and in dealing with flotation<br />

circuits. The vertical sump pump may be supplied as:<br />

� A stand-alone pump with double suction im<strong>pe</strong>ller (Figure 8-37) to be installed in a<br />

concrete or metal sump, particularly with flotation columns<br />

bearing<br />

cartridge<br />

pump<br />

frame<br />

clamping<br />

bolts<br />

for<br />

bearing<br />

cartridge<br />

adjustment<br />

bolt<br />

for<br />

bearing<br />

cartridge<br />

FIGURE 8-34 Adjustment of the pump im<strong>pe</strong>ller by a s<strong>pe</strong>cial bolt between the bearing cartridge<br />

and the frame.


8.54 CHAPTER EIGHT<br />

clamping bolts<br />

bearing cartridge<br />

push bolts<br />

pump frame<br />

FIGURE 8-35 Bearing assembly of the ZJ <strong>slurry</strong> pump (made in China) with adjusting push<br />

and pull bolts. (Courtesy of AJP Services Inc. The distributor for Canada.)<br />

FIGURE 8-36 Sump pump with double suction im<strong>pe</strong>ller. (Courtesy of Mazdak International<br />

Inc.)


(a)<br />

(b)<br />

Rubber-Lined, Acid-Proof Pumps with Double Suction Im<strong>pe</strong>llers<br />

Pump<br />

Size Frame Units A B C D E F G H J K L N P<br />

SP2 BV inch 26 11 32 36 12 20 6 16 20 20 20 40 8<br />

mm 660 280 813 915 305 508 152 406 508 508 508 1016 200<br />

SP3 CV inch 32 14 36 48 16 20 8 22 22 26 26 60 12.6<br />

mm 813 356 915 1220 406 508 203 559 559 660 660 1524 320<br />

SP4 DV inch 37 14 48 60 20 26 10 26 26 30 30 72 14.8<br />

mm 940 356 1220 1524 508 660 254 660 660 762 762 1829 376<br />

SP6 EV inch 48 14 60 72 24 35 14 34 34 38 38 84 19.7<br />

mm 1220 356 1524 1829 610 889 356 864 864 965 965 2134 500<br />

SP8 FV inch 52 14 60 84 14 35 16 47 47 52 52 96 23<br />

mm 1321 356 1524 2987 356 889 406 1194 1194 1321 1321 2438 584<br />

*D is the standard depth—other shaft length are available in 12� increments—consult the plant for critical<br />

s<strong>pe</strong>ed<br />

*E is the minimum priming level<br />

*C and N are typical sump dimensions for the sump<br />

FIGURE 8-37 Dimensions for sump pump and corresponding sump. (Courtesy of Mazdak<br />

International Inc.)<br />

8.55


8.56 CHAPTER EIGHT<br />

� A single suction im<strong>pe</strong>ller with an auger or agitator below the im<strong>pe</strong>ller to agitate settled<br />

solids in floor sumps (Figure 8-38)<br />

� A top suction pump supplied integrally with a metal conical tank, called a “tank pump”<br />

(Figure 8-39)<br />

The vertical <strong>slurry</strong> pump is designed to have all its bearings above the baseplate so as to<br />

be well protected from <strong>slurry</strong> ingestion (Figure 8-40). Due to the depth of the sump, the<br />

design engineer must pay particular attention to the critical s<strong>pe</strong>ed of the pump. For this<br />

reason, the shaft of these pumps can be as large as 200 mm (8�) to offer the necessary<br />

rigidity.<br />

Vertical <strong>slurry</strong> pumps are particularly popular in froth handling circuits. To handle<br />

the combination of solids, air, and liquids, a double suction im<strong>pe</strong>ller is often recom-<br />

wearplate<br />

casing<br />

im<strong>pe</strong>ller<br />

screen<br />

column<br />

agitator<br />

shaft<br />

motor & pulleys<br />

bearing<br />

assembly<br />

baseplate<br />

2" discharge<br />

eyebolt<br />

fig 8-38<br />

FIGURE 8-38 Sump pump with single suction im<strong>pe</strong>ller and auger to agitate settled solids<br />

particularly suited for mill discharge floor. (Courtesy of Mazdak International Inc.)


tank<br />

shaft<br />

im<strong>pe</strong>ller<br />

casing<br />

THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

motor & pulleys<br />

baseplate<br />

inlet<br />

bearing<br />

assembly<br />

eyebolt<br />

discharge<br />

8.57<br />

39 FIGURE 8-39 Tank sump pump with top suction im<strong>pe</strong>ller and integral tank for particularly<br />

difficult frothy slurries.<br />

mended with vertical sump pumps. As an alternative, tank pumps with top suction (Figure<br />

8-39) are used. In either configuration, the im<strong>pe</strong>ller must be designed to resist air<br />

biting.<br />

A s<strong>pe</strong>cial ty<strong>pe</strong> of process used to extract gold is based on cyanide leaching. Leached<br />

gold is then separated by adsorption, the pro<strong>pe</strong>rty of certain materials such as carbon to<br />

fix gold on their surface. Carbon spheres are used as an adsorption material. This process<br />

is done in s<strong>pe</strong>cial “carbon in leach” or “carbon in pulp” circuits with mixing tanks. The<br />

transfer of these solutions requires recessed or vortex im<strong>pe</strong>llers that can pump without<br />

breaking the carbon lumps. The im<strong>pe</strong>ller is recessed out of the flow as shown in Figure<br />

8-41.<br />

A design that is gaining popularity in plants for recycling newspa<strong>pe</strong>r is the vertical<br />

pump with a recessed im<strong>pe</strong>ller and a chop<strong>pe</strong>r blade. It is not uncommon that the recycling<br />

bins for pa<strong>pe</strong>r now found in every suburb of North America end up containing milk cartons,<br />

plastic bottles, toys, and even pieces of wood. These materials are not very good for<br />

conventional stainless steel pumps with mechanical seals. The cantilever sump pump<br />

(Figure 8-41) with the chop<strong>pe</strong>r offers the ability to pump long fibers while chopping them<br />

and eliminating the maintenance problems of mechanical seals.


FIGURE 8-40 Components of a vertical <strong>slurry</strong> pump showing that the bearings are above<br />

the baseplate. 1. Shaft sea. 2. Top bearing cover. 3. Top bearing. 4. Bearing assembly. 5.<br />

Crease nipple. 6. Bearing locknut. 7. Bearing washer. 8. Bottom bearing—spherical roller for<br />

heavy duty. 9. Discharge pi<strong>pe</strong>. 10. Baseplate. 11. Shaft seal. 12. Bottom hub. 13. Pedestal—<br />

O<strong>pe</strong>n structure. 14. Top suction strainer. 15. Wear plate. 16. Shaft. 17. Shaft sleeve. 18. Double<br />

suction im<strong>pe</strong>ller for minimum thrust loads. 19. Pump casing. 20. Lower suction strainer.<br />

(Courtesy of Mazdak International Inc.)<br />

8.58


THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

8-8 GRAVEL AND DREDGE PUMPS<br />

Wet end with cutter<br />

stator<br />

Rotor<br />

8.59<br />

FIGURE 8-41 Vertical <strong>slurry</strong> pump with a recessed im<strong>pe</strong>ller. This pump is suitable for carbon<br />

transfer in gold cyanide circuits and for wastewater treatment applications. The addition of<br />

a cutter (rotor and stator) renders this pump particularly suitable for certain applications such<br />

as sewage treatment and newspa<strong>pe</strong>r recycling plants.<br />

Hard metal pumps play an important role in dredging lakes and ports. Some sizes are<br />

presented in Table 8-12. A typical construction of a dredge pump is presented in Figure<br />

8-42. Dredge pumps are designed to handle particularly large boulders and lumps of<br />

clay. Some of the largest dredge pumps are designed to handle 6.3 m 3 /s or 100,000 US<br />

gpm.<br />

A s<strong>pe</strong>cial low-pressure, high-flow pump called the ladder pump is designed to be<br />

mounted at the tip of the suction arm. Its purpose is essentially to move the material up to<br />

the boat hop<strong>pe</strong>r or up to a booster pump on a hop<strong>pe</strong>r. The booster pump is designed for<br />

higher discharge head.<br />

A particular ty<strong>pe</strong> of pump is the phosphate-matrix-handling pump. It does resemble in<br />

many as<strong>pe</strong>cts a sort of dredge pump, but is built of materials to handle both corrosion and


8.60 CHAPTER EIGHT<br />

Precision<br />

machined<br />

heavy<br />

duty shell<br />

Adjusting<br />

bolt allows<br />

for easy<br />

adjustment of<br />

im<strong>pe</strong>ller for<br />

pro<strong>pe</strong>r<br />

o<strong>pe</strong>rating<br />

clearances<br />

Integral stuffing box<br />

shell mount<br />

Separate<br />

thrust bearing<br />

wear. It is often driven by a diesel engine through a gearbox. The complete baseplate with<br />

driver and pump are relocated from one area to another as mining is done.<br />

8-9 AFFINITY LAWS<br />

Heavy duty bearing<br />

assembly for high power<br />

& loading conditions<br />

FIGURE 8-42 Components of the Marathon dredge pump. (Courtesy of Mobile Pulley and<br />

Machine Works.)<br />

Affinity laws are used to predict the effects of changing the s<strong>pe</strong>ed of a pump, trimming an<br />

im<strong>pe</strong>ller, and extrapolating the <strong>pe</strong>rformance of a pump from case (A) to case (B). They<br />

state that:<br />

2 2 HA/HB = N A/N B (8-63)<br />

H A/H B = D A 2 /DB 2 (8-64)<br />

H A/H B = N A 2 /N B 2 (8-65)<br />

Q A/Q B = D A/D B<br />

Q A/Q B = N A/N B<br />

One piece shell and engine<br />

side door/liner for minimum<br />

parts replacement<br />

Solid im<strong>pe</strong>ller hub eliminates problems<br />

with threaded or bolted inserts<br />

Advanced design im<strong>pe</strong>ller. Good<br />

hydraulic <strong>pe</strong>rformance without<br />

sacrificing spherical clearance<br />

OPTIONAL ONE PIECE DOOR/SIDE<br />

LINER DESIGN AVAILABLE<br />

(8-66)<br />

(8-67)


THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

8-10 PERFORMANCE CORRECTIONS FOR<br />

SLURRY PUMPS<br />

Slurry pumps are designed and tested for hydraulics using water as a reference fluid.<br />

However, they are designed to handle large spherical rocks. Often, four-vane im<strong>pe</strong>llers<br />

are less efficient but outlast other ty<strong>pe</strong>s, particularly on abrasive slurries. Attempts have<br />

been made to quantify and qualify the spacing of vanes on the <strong>pe</strong>rformance of dredge<br />

<strong>slurry</strong> pumps. As early as 1932, Fischer and Thoma conducted tests on a pump built of<br />

a transparent material. Although the flow was observed to be close to the designed value<br />

at best efficiency point, it quickly deviated at other values. They observed a large<br />

area of flow separation on the trailing edge of the vane, with reverse flow in certain instances.<br />

Understanding the effects of solids on centrifugal pumps has been a slow process.<br />

Fairbanks (1941) develo<strong>pe</strong>d a theory to correlate the head develo<strong>pe</strong>d by a pump for a <strong>slurry</strong><br />

mixture with the volumetric concentration and s<strong>pe</strong>cific gravity of the solids. He explained<br />

that the fundamental Euler equation could be modified to account for the density<br />

and flow rate of the mixture as:<br />

T = �mQm(r2Vt2m – r1Vt1m) The power needed to pump the mixture by an ideal pump (at 100% hydraulic efficiency)<br />

is then expressed as:<br />

T� = �mQmgHm where � is the angular s<strong>pe</strong>ed . The mixture head is then expressed as two components for<br />

solids and carrier fluid:<br />

Hm = (�/g�m) · [�s · Cv · (r2Vt2s – r1Vt1s) + (1 – Cv) · (r2Vt2L – r1Vt1L) (8-68)<br />

where<br />

Cv = volumetric concentration of solids<br />

�m = density of mixture<br />

�s = density of solids<br />

Fairbanks concluded from his tests on a single pump that :<br />

� The drop in the head-capacity curve varies not only as the concentration increases, but<br />

also as the particle size of the material in sus<strong>pe</strong>nsion increases.<br />

� The fall velocity of the sus<strong>pe</strong>nded material is the most important parameter for predicting<br />

the effect of solids on pump <strong>pe</strong>rformance.<br />

� The power input is a linear relationship of the apparent s<strong>pe</strong>cific gravity of the solids in<br />

sus<strong>pe</strong>nsion<br />

8-10-1 Corrections for Viscosity and Slip<br />

8.61<br />

Viscosity must be taken in account when pumping viscous slurries. Viscosity reduces the<br />

efficiency of pumping and the head develo<strong>pe</strong>d by a pump (Figure 8-43). The Hydraulic<br />

Institute Standards provides correction curves for viscous fluids pumping, but warns<br />

against extrapolating to other pumps or fluids. The Institute does not publish curves for<br />

viscous slurries.<br />

Duchham and Aboutaleb (1976) derived equations to predict the effects of viscosity


8.62 CHAPTER EIGHT<br />

N<br />

H/H<br />

1.4<br />

1.2<br />

Head (Viscous fluid)<br />

Head (Water)<br />

1.0 1.0<br />

0.8 0.8<br />

0.6 0.6<br />

Efficiency (water)<br />

0.4 0.4<br />

Efficiency<br />

(viscous fluid)<br />

0.2 0.2<br />

0.0 0.0<br />

0.0 0.5 1.0 1.5<br />

Fig 843<br />

Q/Q<br />

N<br />

FIGURE 8-43 Effect of viscosity on the <strong>pe</strong>rformance of centrifugal pumps.<br />

and density on the flow rate, head, and power consumption by comparing particle<br />

Reynolds number and power factor. Their analysis did not present a definite appreciation<br />

of the effects of viscosity.<br />

Sheth et al. (1987) investigated slip factors for <strong>slurry</strong> pumps by conducting tests on a<br />

Wilfley pump. The pump had a 267 mm (10.5 in) diameter, 27 mm (1.06 in) blade width,<br />

and a discharge angle of 31°. The following equation was derived by Sheth et al. (1987)<br />

to account for the effects of the <strong>slurry</strong> mixture carrier densities:<br />

0.12<br />

�s� � = 0.0989 – 0.00157 � � 0.5 ND2 �<br />

�m ��<br />

� �<br />

�L · N · D �L Q<br />

2<br />

where<br />

� s = slip factor<br />

� = dynamic (absolute viscosity) of liquid carrier<br />

D imp = im<strong>pe</strong>ller diameter<br />

N = rotating s<strong>pe</strong>ed of im<strong>pe</strong>ller<br />

� m = density of <strong>slurry</strong> mixture<br />

� L = density of liquid carrier<br />

Q = flow rate<br />

N<br />

(8-69)


The above equation is empirical and the exponents and coefficients may change for different<br />

pump designs. More research work on different designs would have to be published<br />

before a universal formula is adopted.<br />

Example 8-6<br />

A <strong>slurry</strong> pump is to be designed to pump <strong>slurry</strong> under the following conditions:<br />

maximum s<strong>pe</strong>ed at intake 4 m/s (13 ft/s)<br />

flow rate 120 L/s (1858 USGPM)<br />

head 40 m (131 ft)<br />

<strong>slurry</strong> density 1470 kg/m 30 (SG m = 1.47)<br />

water carrier<br />

<strong>slurry</strong> viscosity 100 mPa · s<br />

max solid particle size 25 mm (1 in)<br />

Using the Sheth formula, determine the geometry of the im<strong>pe</strong>ller.<br />

Solution<br />

suction area = = 0.04 m2 suction diameter = 0.225 m (8.85 in)<br />

suction area at 4 m/s = 0.03 m2 0.120 m<br />

suction diameter = 0.195m (7.7 in)<br />

3 /s<br />

��<br />

3<br />

Or, calling a = N · D 2 :<br />

If A = 30, then:<br />

If a = 40, then:<br />

If a = 20, then:<br />

If a = 15, then:<br />

If a = 10, then:<br />

THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

0.12<br />

�s� � = 0.0989 – 0.00157 � � 0.5 ND2 �<br />

�m ��<br />

� �<br />

�L · N · D �L Q<br />

2<br />

� s · 0.331 = 0.0989a 0.12 – 0.0067a 0.62<br />

� s = 0.298a 0.12 – 0.020a 0.62<br />

� s = 0.298 × 1.5 – 0.0202 × 8.23<br />

� s = 0.28<br />

� s = 0.464 – 0.198 = 0.265<br />

� s = 0.427 – 0.129 = 0.297<br />

� s = 0.4124 – 0.108 = 0.304<br />

� s = 0.393 – 0.084 = 0.31<br />

8.63


8.64 CHAPTER EIGHT<br />

Let us assume a = 10, then:<br />

10 = N · D 2<br />

Since the particle size passage is 25 mm (� 1�), assume discharge width = 30 mm or<br />

0.03 m (1.18 in). The head ratio � = 2gH/u 2 (in the United States) is:<br />

� = 2�H�[1 – (cm/u2) cot �2] If we assume D = 0.4m (� 16u ), then:<br />

10 = N × 0.16 ⇒ N = 62.5 rev/s<br />

U = 78.5 m/s<br />

Cm = Q/ADis ⇒ discharge area = � × 0.4??(1 – zt/sin �2 )<br />

If b = 30 mm (1.181�) then:<br />

A = � × 0.4 × 0.03 (1 – zt/sin �2 )<br />

= 0.037 (1 – zt/sin �2 )<br />

If Z = 4 vanes and t = 30 mm then:<br />

A = 0.037 (1 – 0.12/sin �2 )<br />

If �2 = 15° then:<br />

A = 0.0198 m2 Cm = 0.12/0.0198 = 6.05 m/s<br />

Cm/U2 = 0.077<br />

� = 2� H� s(1 – (C m/U 2) cot � 2) = 0.44� H<br />

2gH 2 × 9.81 × 40<br />

� = = �� = 0.127<br />

� U 2 2<br />

78.5 2<br />

0.127 = 0.44 �H ⇒ �H = 0.289<br />

This is not a very efficient pump due to the combination of viscosity and solid density:<br />

consumed power = g�QH/�H = 9.81 × 470 × 0.120 × 40/0.289<br />

� 240 kw or 327 hp<br />

It is recommended to install a 400 hp motor.<br />

8-10-2 Concepts of Head Ratio and Efficiency Ratio When<br />

Pumping Solids<br />

Stepanoff (1969) explained that when pumping solids in sus<strong>pe</strong>nsion, a pump im<strong>pe</strong>ller imparts<br />

energy to the carrier liquid. For a homogeneous mixture, he explained, the im<strong>pe</strong>ller<br />

will be able to impart as many feet of mixture as it would have been able to impart head of<br />

water. The <strong>pe</strong>rformance of the im<strong>pe</strong>ller is not impaired but the power consumption increases<br />

linearly with the s<strong>pe</strong>cific gravity of the mixture. In reality, at best efficiency point,<br />

the presence of solids tends to reduce the hydraulic head by the energy wasted to move<br />

them through the im<strong>pe</strong>ller passageways. Similarly, the efficiency of the pump when handling<br />

the mixture will be reduced by the presence of solids. Two factors can be defined<br />

head ratio and efficiency ratio.


THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

8.65<br />

The head ratio is<br />

HR = Hm/Hw (8-70)<br />

or the ratio of head develo<strong>pe</strong>d when pumping <strong>slurry</strong> to the head develo<strong>pe</strong>d when pumping<br />

water.<br />

The efficiency ratio is<br />

ER = Em/Ew (8-71)<br />

or the ratio of efficiency develo<strong>pe</strong>d when pumping <strong>slurry</strong> to the efficiency develo<strong>pe</strong>d<br />

when pumping water.<br />

Stepanoff (1969) indicated that at best efficiency:<br />

HR = ER Stepanoff (1969) reported work by Japanese investigators who indicated that tests on carbide<br />

slurries tended to show that the head–capacity ratio may increase or decrease de<strong>pe</strong>nding<br />

on whether the solids concentration tended to cause the <strong>slurry</strong> to behave as a<br />

Newtonian or non-Newtonian mixture.<br />

Reviewing published data between 1941 and 1971, Hunt and Faddick (1971) reported<br />

that various tests in different labs and field applications confirmed that:<br />

� The drop in head (in feet of mixture flowing) develo<strong>pe</strong>d for a given volumetric discharge<br />

rate decreased as the concentration of the solids in sus<strong>pe</strong>nsion increased.<br />

� The required brake horsepower for a given pump o<strong>pe</strong>rating at a given capacity increased<br />

as the concentration of solid material in sus<strong>pe</strong>nsion increased.<br />

� The efficiency at a given capacity decreased as the concentration of the solid material<br />

in sus<strong>pe</strong>nsion increased.<br />

Hunt and Faddick (1971) simulated the <strong>pe</strong>rformance of centrifugal pumps pumping wood<br />

chips by tests using rectangular plastic parts with an average s<strong>pe</strong>cific gravity of 1.02.<br />

They used four different im<strong>pe</strong>ller designs in two different volute designs. There was no<br />

consistency in the extent of head drop or efficiency with solid concentration, and the results<br />

indicated that the actual design of the pump was a very important factor. A difference<br />

of head and efficiency of 5 to 7% was noticed for the different designs. The authors<br />

therefore discouraged applying head and efficiency ratio factors for one pump to another<br />

pump of a different geometry, but encouraged further research into the mechanisms of<br />

flow through the rotating passages of these pumps.<br />

It is important to appreciate the work of Hunt and Faddick. Often, a pump vendor will<br />

produce a chart or curve to obtain the head and efficiency ratio. The limitations of such<br />

curves are that they apply only to pumps of similar geometrical design. The discrepancy<br />

of 5–10% between one design and another may have to be absorbed by the motor.<br />

McElvain (1974) published data on the effects of solids on pump <strong>pe</strong>rformance. He<br />

worked on the concept of the head and efficiency reduction factors defined as:<br />

RH = 1 – HR (8-72)<br />

R� = 1 – ER (8-73)<br />

He tested im<strong>pe</strong>llers up to a diameter of 35 cm (13.78 in) and on various concentrations of<br />

silica and one grade of heavy mineral. He develo<strong>pe</strong>d a set of curves and established a relationship<br />

between volumetric concentration and the head and efficiency reduction factors<br />

as:<br />

RH = R� = 5 · K · CV (8-74)


8.66 CHAPTER EIGHT<br />

N<br />

H/H<br />

1.2<br />

Head (Slurry)<br />

Head (Water)<br />

1.0 1.0<br />

0.8 0.8<br />

0.6 0.6<br />

Efficiency(water)<br />

0.4 0.4<br />

Efficiency (<strong>slurry</strong>)<br />

0.2 0.2<br />

0.0 0.0<br />

0.0 0.5 1.0 1.5<br />

Fig 8-44<br />

Q/Q<br />

N<br />

FIGURE 8-44 Effect of solids on the <strong>pe</strong>rformance of centrifugal pumps.<br />

The K factor was then plotted against the d 50 and for solids of various s<strong>pe</strong>cific gravity<br />

(see Figure 8-45). The assumption that R H = R � was accepted to hold true for <strong>slurry</strong> volumetric<br />

concentrations smaller than 20%. This covers a substantial number of pump applications.<br />

Example 8-7<br />

Heavy metal oxide <strong>slurry</strong> is to be pum<strong>pe</strong>d at a volumetric concentration of 18%. The s<strong>pe</strong>cific<br />

gravity of the solids is 5.0 and the d50 is 400 �m. The calculated head on <strong>slurry</strong> is 35<br />

m. Determine the head ratio and the equivalent water head on the pump <strong>pe</strong>rformance curve.<br />

Solution<br />

Using the McElvain equation, the value K is determined from the lower curve at 0.38.<br />

Substituting in Equation 8-74, at a volumetric concentration of 18% (less than 20%):<br />

RH = R� = 5 · K · CV = 5 · 0.38 · 0.18 = 0.342<br />

Substituting into Equation 8-72:<br />

HR = 1 – RH = 0.658<br />

Since the calculated head for friction, the equivalent value on water is<br />

Hw = Hm/HR = 35 m/0.658 = 53.2 m<br />

The engineer must therefore select the appropriate pump s<strong>pe</strong>ed from the pump curve that<br />

would develop 53.2 m.<br />

N


K-Factor<br />

0.1<br />

0.2<br />

0.3<br />

0.4<br />

0.5<br />

Sellgren and Vappling (1986) reported that at high volumetric concentration the efficiency<br />

ratio was smaller than the head ratio, thus indicating a more pronounced loss of efficiency.<br />

Sellgen and Addie (1993) reported losses as low as half of those predicted by McElvain.<br />

The curves of McElvain do not take into account another important factor, namely<br />

the ratio of particle size to im<strong>pe</strong>ller diameter. Burgess and Reizes (1976) proposed that<br />

the head ratio and efficiency ratio were a function of three parameters:<br />

1. Weight concentration<br />

2. Ratio of d 50 to im<strong>pe</strong>ller diameter<br />

3. S<strong>pe</strong>cific gravity of the solid particles<br />

THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

10 100<br />

1000<br />

S = 5.0<br />

s<br />

Particle Size d ( m)<br />

8.67<br />

Sellgen and Addie (1993) indicated that there is a size effect and that head and efficiency<br />

losses were less drastic in large pumps than in small pumps (Figure 8-46). This clearly<br />

demonstrates the importance of pump design on <strong>pe</strong>rformance.<br />

The importance of pump design on the head and efficiency ratio was confirmed by<br />

Czarnota et al. (1996) through tests on ITT-Flygt submersible pumps. Their work confirmed<br />

that high-efficiency pumps suffered from less degradation of <strong>pe</strong>rformance than<br />

less efficient pumps. Head reduction was confirmed to be a linear function of the volumetric<br />

concentration of solids. Larger particles were found to slip more than smaller particles.<br />

An important factor they reported is that settling or separation can occur due to<br />

centrifugal forces. These forces are proportional to the square of the radius, and in the<br />

presence of large particles can lead to partial blockage, higher water velocity, and more<br />

slip between solids and liquid. A well-mixed particle distribution tended to decrease derating<br />

of pumps.<br />

Russian engineers develo<strong>pe</strong>d a very advanced mathematical model based on full<br />

screen analysis instead of the average d 50, which has been the focus of most equations in<br />

50<br />

2.65<br />

4.0<br />

1.5<br />

10000<br />

FIGURE 8-45 Correction to the head K factor of centrifugal pumps on the basis of the s<strong>pe</strong>cific<br />

gravity and particle size S s = s<strong>pe</strong>cific gravity of solids. (After McElvain, 1974.)


8.68 CHAPTER EIGHT<br />

FIGURE 8-46 Effect of the size of the pump im<strong>pe</strong>ller on the correction factor for head R H<br />

for <strong>slurry</strong> at a weight concentration of 42%. (From Sellgren and Addie, 1992.).<br />

Australia, Euro<strong>pe</strong>, and North America. The work of Kuznetsov and Samoilovich (1985,<br />

1986) was summarized by Angle et al. (1997). These advanced mathematical models <strong>pe</strong>rmit<br />

corrections based on the number of vanes, discharge angle of the vanes, and volumetric<br />

concentration of each range of diameter of solids in the <strong>slurry</strong>. It would be very appropriate<br />

to explore these models; however, when examining a worn-out im<strong>pe</strong>ller, as in<br />

Figure 8-47, the reader may wonder how practical such models may be.<br />

8-10-3 Concepts of Head Ratio and Efficiency Ratio Due to<br />

Pumping Froth<br />

Flotation froth is complex <strong>slurry</strong> and may contain an important amount of air and gases<br />

(Figure 8-48). The industry uses the froth factor as a measure. Basically, it is determined<br />

by filling a column of flotation <strong>slurry</strong> and measuring the height H0. It is then left for 24 hr<br />

to rest. The height of the <strong>slurry</strong> H� is then measured. The froth factor F is defined as:<br />

F = H0/H� (8-75)<br />

Using this concept of froth factor to size pumps must be done very carefully. Different<br />

grades of froth leads to different levels of entrained gases, as shown in Table 8-15. Conventional<br />

centrifugal pumps can not handle excessive amounts of entrained gases.<br />

A very common misunderstanding in the industry is that the flow rate of <strong>slurry</strong> must<br />

be multiplied by the froth factor to size the pump. This violates a very fundamental principle<br />

that gases or air are compressible fluids. In other words, as the bubbles pass through<br />

the im<strong>pe</strong>ller they are compressed and reduced in size. In fact, the pro<strong>pe</strong>r sizing of <strong>slurry</strong><br />

pumps to handle froth must be based on a full examination of the system. For example, if


THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

8.69<br />

FIGURE 8-47 Worn-out im<strong>pe</strong>ller, showing gradual degradation of the im<strong>pe</strong>ller tip diameter<br />

and vanes. Deterioration of hydraulics and head efficiency may occur throughout the wear life<br />

of the im<strong>pe</strong>ller and pump.<br />

FIGURE 8-48 Flotation <strong>slurry</strong> froth contains sufficient air bubbles to degrade the <strong>pe</strong>rformance<br />

of centrifugal pumps. (Courtesy of EOMCO Process Equipment Co.)


8.70 CHAPTER EIGHT<br />

TABLE 8-15 Correlation between Froth Factor and Percentage of Entrained Air<br />

Froth factor % entrained air Example<br />

1.5 2–3% Normal flotation tailings<br />

2.0 3–5% Flotation tailings with minimum retention time<br />

2.5 5–7% Tenacious flotation tailings with minimum retention time<br />

3.0 > 7% Froth with very fine particles<br />

the flotation cells are away from the pump and the froth is transported by gravity in launders,<br />

it may be argued that the surface area of the launders acts as a deaerator for air removal.<br />

In that case, does a 24 hr tube test apply well?<br />

The correct approach is in fact to remove as much of the air as possible before the<br />

froth enters the pump feed sump. This sump may also be designed in a conical sha<strong>pe</strong> to<br />

maximum surface area at the top.<br />

Certain forms of froth are very difficult to pump, such as the ty<strong>pe</strong> associated with tar<br />

sands, in which viscosity plays a major role. Efficiency as low as 10% was reported with<br />

conventional pumps.<br />

Cap<strong>pe</strong>lino et al. (1992) presented a very thorough study on the <strong>pe</strong>rformance of centrifugal<br />

pumps with o<strong>pe</strong>n im<strong>pe</strong>llers with emphasis on pulp and pa<strong>pe</strong>r flotation circuits<br />

and deinking cells. High-consistency stock (12%) can have as much as 20–28% entrained<br />

air.<br />

At the inlet to the im<strong>pe</strong>ller, the pressure drop tends to cause an expansion of the air<br />

and gases, and this indicates well that the concept of the froth factor can be misleading<br />

Example 8-8<br />

The height of the liquid in a froth cell is 30 m above the pump. The depression at the inlet<br />

to the pump is about 6 m. Determine the expansion of gases, assuming a barometric pressure<br />

of atmospheric air at 9.5 m. The pump is designed to deliver a total head of 54 m.<br />

Determine the final volume of the gases.<br />

Solution<br />

The effective absolute head in the sump is:<br />

30 + 9.5 = 39.5 m<br />

Due to the depression of 6 m, the absolute pressure is then:<br />

39.5 – 6 = 33.5 m<br />

The expansion ratio at constant tem<strong>pe</strong>rature is:<br />

39.5/33.5 = 1.179<br />

The absolute discharge head is:<br />

suction head + TDH + atmospheric barometric height = 30 + 54 + 9.5 = 93.5 m<br />

Ratio of discharge to suction absolute head is:<br />

93.5/39.5 = 2.37<br />

The size of the air or gas bubbles will then shrink by the inverse of this ratio, or 42.2%.<br />

Since the laws of thermodynamics apply, the concept of a constant froth factor is illusive.<br />

It would be a grave error to size piping and equipment based on the suction froth<br />

factor.


The <strong>pe</strong>rformance of <strong>slurry</strong> pumps deteriorates in the presence of entrained gases. Cap<strong>pe</strong>lino<br />

et al. (1992) have therefore proposed to define appropriate head and power correction<br />

factors as:<br />

or<br />

THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

8.71<br />

FIGURE 8-49 Head and power correction factors for entrained gas due to flotation circuits.<br />

(From Cap<strong>pe</strong>llino, Roll, and Wilson, 1992. Reproduced by <strong>pe</strong>rmission of Texas A&M University.)<br />

head measured with entrained gas<br />

HF = ����<br />

(8-76)<br />

head measured without entrained gas<br />

power measured with entrained gas<br />

PF = ����<br />

(8-77)<br />

power measured without entrained gas<br />

S<strong>pe</strong>cial pumps are available for handling froth and entrained gases. One interesting<br />

design is the Sulzer–Ahlstrom ART pump. It features holes through the im<strong>pe</strong>ller leading


8.72 CHAPTER EIGHT<br />

straight to an ex<strong>pe</strong>ller at the back of the im<strong>pe</strong>ller. This ex<strong>pe</strong>ller discharges the air through<br />

a separate discharge flange at the back of the casing. In some other designs, the stuffing<br />

box is connected to an external liquid ring vacuum pump that can remove any entrained<br />

air in the <strong>slurry</strong>. The manufacturers of pulp and pa<strong>pe</strong>r pumps have also designed s<strong>pe</strong>cial<br />

im<strong>pe</strong>ller with protruding vanes that extend into the suction pi<strong>pe</strong> to break up any large air<br />

particles. This concept is gaining popularity in some oil sand applications to handle particularly<br />

thixotropic and viscous froth.<br />

Dredge pumps sometimes face a similar problem. Gases are disturbed or released (particularly<br />

methane) during certain phases of dredging and end up accumulating in the ladder<br />

pump. Herbich and Miller (1970) conducted extensive test work on the effect of air on<br />

the development of head. Herbich (1992) proposed that s<strong>pe</strong>cial air removal <strong>systems</strong> be installed<br />

on the suction side of the pump with an ejector, as this is better than a vacuum<br />

pump.<br />

8-11 CONCLUSION<br />

In this chapter, some of the important parameters that give <strong>slurry</strong> pumps their final sha<strong>pe</strong><br />

were examined. It is obvious that <strong>slurry</strong> pumps are different from water pumps and that<br />

considerable research should be undertaken in fields such as pump-out vanes, ex<strong>pe</strong>ller design,<br />

and effects of wide im<strong>pe</strong>llers on <strong>pe</strong>rformance.<br />

The successful <strong>pe</strong>rformance of these pumps de<strong>pe</strong>nds on their resistance to wear. The<br />

<strong>slurry</strong>—in terms of its composition and concentration, and in terms of any froth-induced<br />

gases—is the determining factor for power consumption and the final hydraulics across<br />

the im<strong>pe</strong>ller and casing. These parameters are extremely important for the successful installation<br />

of these pumps.<br />

8-12 NOMENCLATURE<br />

A Factor to calculate slip, de<strong>pe</strong>nding on the use of volute or diffuser<br />

Ap Equivalent casing area for stress calculations<br />

bt Thread width at the root<br />

b2 Width of the im<strong>pe</strong>ller at the im<strong>pe</strong>ller tip diameter<br />

B2 Width of the casing at the im<strong>pe</strong>ller tip diameter<br />

BHP Power in bhp<br />

C Constant<br />

Cm Meridional velocity across the im<strong>pe</strong>ller<br />

Cv volumetric concentration of solids<br />

Cw Concentration by weight of the solid particles in <strong>pe</strong>rcent<br />

Cp Heat capacity<br />

D Equivalent diameter of casing<br />

d2 The tip diameter of the im<strong>pe</strong>ller vanes<br />

dSL Diameter of shaft sleeve<br />

dm Pitch diameter<br />

ER Efficiency ratio<br />

Fth Force due to thread pull<br />

FA Axial thrust on shaft due to hydraulic forces<br />

Force on casing due to design pressure<br />

F p


Fp Force due to belts at the drive end of the pump<br />

FR Radial thrust<br />

G Acceleration due to gravity (9.78 to 9.81 m/s2 or 32.2 ft/sec)<br />

H Height of the thread tooth = (major diameter – minor diameter)/2; in the case of<br />

ACME threads h = p/2<br />

H1 Head at the medium diameter of the eye of the im<strong>pe</strong>ller<br />

H2 Head at the tip diameter of vane of the im<strong>pe</strong>ller<br />

H30 Head at 30% of best efficiency capacity<br />

HE Euler ideal head for an im<strong>pe</strong>ller<br />

HIMP SV Shut-off head due to the im<strong>pe</strong>ller<br />

HR Head ratio<br />

HVOL SV Shut-off head due to the volute<br />

HSV Total shut-off head<br />

HV Vapor head<br />

K Correction factor for the head ratio<br />

Kx Ky L<br />

Coefficient to determine Xv Coefficient to determine Yv length of a full turn in a shaft thread = pitch for single start threads<br />

Lth The length of the shaft thread<br />

m Exponent in vortex equation<br />

M Number of pumps in series<br />

N Rotational s<strong>pe</strong>ed of the pump in rev/min<br />

NPSH Net Positive Suction Head<br />

Nq S<strong>pe</strong>cific s<strong>pe</strong>ed in SI units<br />

NUS S<strong>pe</strong>cific s<strong>pe</strong>ed in US units<br />

NSS Suction s<strong>pe</strong>cific s<strong>pe</strong>ed<br />

PA Atmospheric pressure<br />

Pb Pe Ps PD Pressure at the root of the pump-out vanes Rb Pressure at the surface of the liquid in absolute terms on the suction side<br />

Pressure at the suction diameter d1 Pressure losses between the surface of the liquid and the pump, due to friction,<br />

valves, etc.<br />

PW Pump power in Watts<br />

PV Vapor pressure<br />

Q Flow rate<br />

R1 Root radius of the vanes of the im<strong>pe</strong>ller<br />

R2 Tip radius of the vanes of the im<strong>pe</strong>ller<br />

R3 Radius of the smaller circle of a twin circle volute<br />

R4 Radius of the larger circle of a twin circle volute<br />

Rb Radius at the root of the pump out vanes<br />

RC Cutwater radius<br />

RH Head correction factor<br />

R� Efficiency correction factor<br />

RMD Meridional radius of the volute at the throat<br />

Rr Tip radius of the pump-out vanes<br />

Rs Tip radius of the shaft sleeve<br />

Rv0 Local radius of vanes<br />

R1 Radius of the root of the im<strong>pe</strong>ller vane<br />

R2 Tip radius of the vanes of an im<strong>pe</strong>ller<br />

Rv0 Radius of vortex<br />

Reaction force at the drive end bearing<br />

R D<br />

THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

8.73


8.74 CHAPTER EIGHT<br />

RW Reaction force at the wet end bearing<br />

s Gap between the edge of the pump-out vanes and the back wear plate of the casing<br />

S Static moment, obtained by graphical integration along the meridional plane of<br />

the vanes<br />

Sb Bending stress at the root of the shaft thread<br />

Ss Shear stress at the root of the shaft thread<br />

Tq Torque<br />

t Depth of the pump-out vanes<br />

tC Thickness of the pump casing<br />

tL Thickness of the liner<br />

TS Thrust on suction side<br />

TSL Thrust on shaft sleeve<br />

TDH Total Dynamic Head<br />

U Tip s<strong>pe</strong>ed<br />

V Absolute velocity across the im<strong>pe</strong>ller<br />

W Relative velocity across an im<strong>pe</strong>ller<br />

Wimp Weight of im<strong>pe</strong>ller<br />

Wp Weight of pulleys<br />

X Total gap between the pump-out vanes’ im<strong>pe</strong>ller surface and the pump back<br />

plate<br />

XV Width of the volute in the x-direction<br />

YV Width of the volute in the y-direction<br />

Z Number of vanes<br />

Z1 Geodetic elevation of liquid surface above the centerline of the pump im<strong>pe</strong>ller<br />

Geodetic elevation of the centerline of the pump im<strong>pe</strong>ller<br />

Z e<br />

Greek Letters<br />

� Angular inclination of the vane with res<strong>pe</strong>ct to the tangent<br />

� Slip factor<br />

� Density of liquid<br />

� Cavitations parameter<br />

� Angular velocity<br />

�liq Angular velocity of the liquid in the gap between the pump-out vanes and the<br />

pump back plate<br />

�imp Rotational velocity of the im<strong>pe</strong>ller or ex<strong>pe</strong>ller<br />

�SI Head coefficient to SI convention<br />

�US Head coefficient to US convention<br />

Subscripts<br />

1 At the root of the vane<br />

2 At the tip of the vane<br />

E Eye of the im<strong>pe</strong>ller<br />

F Front shroud<br />

R Rear shroud<br />

Imp Im<strong>pe</strong>ller<br />

Liq Liquid<br />

Sl Sleeve


8-13 REFERENCES<br />

THE DESIGN OF CENTRIFUGAL SLURRY PUMPS<br />

8.75<br />

Abulnaga, B. E. 2001. Recommendations for the design of mill discharge <strong>slurry</strong> pumps. Mazdak International<br />

Inc. Internal Report 02/2001 (unpublished).<br />

Addie, G. R. and F. W. Helmly. 1989. Recent improvements in dredge pump efficiencies and suction<br />

<strong>pe</strong>rformances. Europort Dredging Seminar, Central Dredging Association, Delft, Netherlands.<br />

Anderson, H. H. 1938. Mine pumps. J. Mining Soc. Durham, United Kingdom.<br />

Anderson, H. H. 1977. Statistical records of pump and water turbine effectiveness. International Mechanical<br />

Engineers Conference on Scaling for Performance Prediction in Rotodynamic Pumps.<br />

September, pp. 1–6.<br />

Anderson, H. H. 1980. Centrifugal Pumps. Trade and Technical Press: UK.<br />

Anderson, H. H. 1984. The area ratio system. World Pumps, 201.<br />

Angle, T. and J. Crisswell (Editors). 1977. Slurry Pump Manual. Salt Lake City, Utah: Envirotech.<br />

Angle, T. and A. Rudonov. 1999. Slurry Pump Manual. Salt Lake City, Utah: Envirotech.<br />

ANSI/ASME B106.1. 1985. Design of Transmission Shafting. American Society of Mechanical Engineers,<br />

New York.<br />

Burgess, K. E. and B. E. Abulnaga. 1991. The application of finite element methods to Warman<br />

pumps and process Equipment. Pa<strong>pe</strong>r presented to the Fifth International Conference on Finite<br />

Element Analysis in Australia, University of Sydney, Australia, July 1991.<br />

Burgess, K. E. and J. A. Reizes. 1976. The effect of sizing, s<strong>pe</strong>cific gravity and concentration on the<br />

<strong>pe</strong>rformance of centrifugal <strong>slurry</strong> pumps. Proc. Inst. Mech. Eng., 190, 36.<br />

Cap<strong>pe</strong>llino, C. A., D. Roll, and G. Wilson. 1992. Design considerations and application guidelines<br />

for pumping liquids with entrained gas using o<strong>pe</strong>n im<strong>pe</strong>ller centrifugal pumps. Proceedings of<br />

the Ninth International Pump Users Symposium, Texas A&M University.<br />

Czarnota, Z., M. Fahlgren, M. Grainger, and S. Saunders. 1996. The effects of slurries on the <strong>pe</strong>rformance<br />

of submersible pumps. BHR Group Hydrotranport, 13, 643–655.<br />

Duchham C. D. and Y. K. A. Aboutaleb. 1976. Some tests in a single stage semi-o<strong>pe</strong>n im<strong>pe</strong>ller centrifugal<br />

pump handling coal dust slurries. In Proceedings Pumps and Turbine Conferences, Vol<br />

1.<br />

Fairbanks, L. C. Jr. 1941. Effects on the characteristics of centrifugal pumps. Solids in Sus<strong>pe</strong>nsion<br />

Symposium, Proc. Am. Soc. Civ. Eng., 129, 129.<br />

Fischer, K. and D. Thoma. 1932. Investigation of the flow conditions in a centrifugal pump. Transactions<br />

ASME, 54.<br />

Frost, T. H. and E. Nielsen. 1991. Shut-off head of centrifugal pumps and fans. Proc. Inst. Mech.<br />

Eng., 205, 217–223.<br />

Herbrich, J. 1991. Handbook of Dredging Engineering. New York: McGraw-Hill.<br />

Herbich, J. B. and R. J. Christopher. 1963. Use of high s<strong>pe</strong>ed photography to analyze particle motion<br />

in a model dredge pump. In Proceedings of the International Association for Hydraulic Research,<br />

London England.<br />

Herbich, J. B. and R. E. Miller. 1970. Effect of air content on <strong>pe</strong>rformance of a dredge pump. In Proceedings<br />

of the World Dredging Conference, Wodcon 70, Singapore.<br />

Hunt, A. W and R. F. Faddick. 1971. The effects of solids on centrifugal pump characteristics. In Advances<br />

in Solid–Liquid Flow in Pi<strong>pe</strong>s and Its Application, I. Zandi (Ed.), New York: Pergamon<br />

Press.<br />

Jekat, W. K. 1992. Centrifugal pump theory. Section 2.1 in Pump Handbook, J. Karassik et al. (Eds.),<br />

New York: McGraw Hill.<br />

Kuznetsov, O. V. and D. C. Samoilovich. 1986. Increase of Reliability of Slurry Pumps in Service (in<br />

Russian). Moscow: CINTIchimneftemash, ser.XM-4.<br />

McElvain, R. E. 1974. High pressure pumping. Skillings Mining Review, 63, 4, 1–14.<br />

Pfeiderer, C. 1961. Die Kreiselpum<strong>pe</strong>n. Berlin: Springler-Verlag.<br />

Samoilovich, D. C. 1986. Ex<strong>pe</strong>rimental Study of Slurry Pumps Performances (in Russian). Moscow:<br />

CINTIchimneftemash, ser.XM-4.<br />

Sellgren, A. and L. Vappling. 1986. Effects of highly concentrated slurries on the <strong>pe</strong>rformance of<br />

centrifugal pumps. Proceedings of the International Symposium on Slurry Flows, FED Vol 38,<br />

ASME, USA, pp. 143–148.<br />

Sellgren, A. and G. R. Addie. 1992. Effects of solids on the <strong>pe</strong>rformance of centrifugal <strong>slurry</strong> pumps.


8.76 CHAPTER EIGHT<br />

Pa<strong>pe</strong>r presented at the 10th Colloquium: Massenguttransport durch Rohrietungen in Meschede,<br />

Germany, May 20–22.<br />

Sellgren, A. and G. R. Addie. 1993. Solids effect on the characteristics of centrifugal <strong>slurry</strong> pumps.<br />

Pa<strong>pe</strong>r presented at the 12th International Conference on Slurry Handling and Pi<strong>pe</strong>line Transport,<br />

Brugge, Belgium.<br />

Sheth, K. K., G. L. Morrison, and W. W. Peng. 1987. Slip factors of centrifugal <strong>slurry</strong> pumps.<br />

A.S.M.E. Journal of Fluids Engineering, 109, 313–318.<br />

Stepanoff, A. J. 1969. Gravity flow of bulk solids and transportation of solids in sus<strong>pe</strong>nsion. New<br />

York: Wiley.<br />

Stepanoff, A. J. 1993. Centrifugal and Axial Flow Pumps. Melbourne, FL: Krieger.<br />

Sulzer Pumps. 1998. Centrifugal Pump Handbook. New York: Elsevier.<br />

Turton, R. K. 1994. Rotodynamic Pump Design. Cambridge: Cambridge University Press.<br />

K. C. Wilson, G. R. Addie, and R. Clift. 1992. Slurry Transport Using Centrifugal Pumps. London:<br />

Elsevier Applied Sciences.<br />

Wilson, G. 1976. Construction of solids-handling centrifugal pumps. In Pump Handbook, J. Karassik<br />

et al. (Eds.) New York: McGraw Hill.<br />

Worster, R. C. 1963. The flow in volutes and effect on centrifugal pump <strong>pe</strong>rformance. Proc. Inst.<br />

Mech. Eng., 177, 843.<br />

Further Reading<br />

Kazim, K. A. and B. Maiti. 1997. A correlation to predict the <strong>pe</strong>rformance characteristics of centrifugal<br />

pumps handling slurries. In Proceedings of the Institution of Mechanical Engineers. Part A.<br />

Journal of Power and Energy, 211, A2, 147–157.<br />

Cader, T., O. Masbernat, and M. C. Rocco. 1994. Two phase velocity distributions and overall <strong>pe</strong>rformance<br />

of a centrifugal <strong>slurry</strong> pump. Journal of Fluid Engineering, 116, 316–323.<br />

Gandhi, B. K., S. N. Singh, and V. Seshadri. 2000. Improvements in the prediction of <strong>pe</strong>rformance of<br />

centrifugal <strong>slurry</strong> pumps handling slurries. Proceedings of the Institution of Mechanical Engineers.<br />

Part A. Journal of Power and Energy, 214, 5, 473–486.


CHAPTER 6<br />

SLURRY FLOW IN<br />

OPEN CHANNELS AND<br />

DROP BOXES<br />

6-0 INTRODUCTION<br />

The design of mineral processing plants and tailings disposal <strong>systems</strong> often includes<br />

gravity flows in o<strong>pe</strong>n channels. Such flows are often called slack flows. They involve a<br />

free boundary to the atmosphere. In the past, many launders were designed using rules<br />

of thumb; however, the development of large mines requires a rigorous scientific approach.<br />

Most of the published pa<strong>pe</strong>rs on sediment transportation in o<strong>pe</strong>n channels dwell extensively<br />

on the geophysics of canals and rivers. The field of o<strong>pe</strong>n channel hydraulics is<br />

so vast and complex that the reader may have to consult various reference books such<br />

as Graf (1971). One subject of great interest to civil engineers is the carrying capacity<br />

of the channel for sediments. This is often termed the sediment load and is measured as<br />

a function of the flow rate, width of the channel, and sediment concentration. Using<br />

channels to transport solids has limitations due to the fact that no pumps are used to<br />

force the flow.<br />

Many pa<strong>pe</strong>rs have been published over the years on the geophysics of rivers and the<br />

maximum <strong>pe</strong>rmissible s<strong>pe</strong>eds used to avoid scouring and removal of bed materials.<br />

Transferring the knowledge about scouring s<strong>pe</strong>eds of canals and rivers into useful information<br />

for a designer of a hydrotransport system is not a straightforward process. In fact<br />

there, is not a single unified mathematical model to represent <strong>slurry</strong> flows in o<strong>pe</strong>n channels.<br />

In this chapter, a methodology is presented to estimate the friction losses for <strong>slurry</strong><br />

flows in o<strong>pe</strong>n channels, cascades, drop boxes, and distribution boxes. In the last twenty<br />

years, new developments in thickeners encouraged various o<strong>pe</strong>rators to develop the concept<br />

of adding flocculants to launders. Tailings and concentrate slurries are thereby allowed<br />

to flow at higher and higher concentrations in gravity modes. The engineer must<br />

take into account the rheology, particularly certain as<strong>pe</strong>cts of high yield stress and non-<br />

Newtonian characteristics.<br />

Mineral processing plants often divide or combine flows in drop boxes, distribution<br />

boxes, and plunge pools. The design principles of such entities are presented at the end of<br />

the chapter.<br />

6.1


6.2 CHAPTER SIX<br />

6-1 FRICTION FOR SINGLE-PHASE FLOWS<br />

IN OPEN CHANNELS<br />

The words flume, launder, o<strong>pe</strong>n channel, and slack flow are often used to express the<br />

same thing. In the following discussion, these words will be used interchangeably. Even<br />

though launders are crucial to mining, very little research on the subject has been published.<br />

Despite the lack of reference material on launders for slurries, it is important to<br />

start from basic principles. The analysis will focus initially on water flows. The reader<br />

will then be introduced to the complexity of <strong>slurry</strong> flows. The reader should appreciate<br />

that an up<strong>pe</strong>r practical limit on these flows is a 65% concentration of solids by weight.<br />

Since the flow does not fill the launder or pi<strong>pe</strong>, the hydraulic diameter is the defined as<br />

the equivalent diameter of flow for an o<strong>pe</strong>n channel. The hydraulic radius is defined as<br />

the ratio of the area of the flow to the wetted <strong>pe</strong>rimeter. It is also called the hydraulic<br />

mean depth in certain Euro<strong>pe</strong>an books.<br />

A<br />

RH = � (6-1)<br />

P<br />

FIGURE 6-1 Large concrete structures offer a method of conveying large quantities of <strong>slurry</strong>.<br />

This structure was built to convey 150,000 tons <strong>pe</strong>r day of soft high clay tailings at a Peruvian<br />

cop<strong>pe</strong>r mine.


And the hydraulic diameter is<br />

4A<br />

DH = � (6-2)<br />

P<br />

Figure 6-2 shows various possible sha<strong>pe</strong>s for o<strong>pe</strong>n launders with methods to estimate<br />

wetted area and hydraulic radius. The most common launders in mining and dredging are,<br />

however, circular and rectangular in sha<strong>pe</strong>. Sometimes a circular pi<strong>pe</strong> is o<strong>pe</strong>ned and vertical<br />

walls are added to produce a U-sha<strong>pe</strong>.<br />

The friction loss for a closed channel and steady-state single-phase flow was examined<br />

in Chapter 2. Using the Darcy factor, the head loss for an o<strong>pe</strong>n launder can be expressed<br />

in terms of the hydraulic radius:<br />

h = (6-3)<br />

Since the Darcy friction coefficient fD is usually accepted as four times the fanning<br />

friction coefficient fN, Equation 6-3 for a <strong>slurry</strong> may be rewritten in terms of the fanning<br />

factor fN, discussed in Chapter 2.<br />

h = (6-4)<br />

For a fully develo<strong>pe</strong>d and uniform flow, the slo<strong>pe</strong> or energy gradient of an o<strong>pe</strong>n launder<br />

is established in terms of the head loss <strong>pe</strong>r unit of length (Henderson, 1990):<br />

fNU S = = (6-5)<br />

2<br />

fNLU H<br />

� �<br />

L 2gRH<br />

2<br />

fDLU �<br />

2gRH<br />

2<br />

�<br />

2g(4RH)<br />

2�<br />

2�<br />

SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

2� < �<br />

A = R 2 (� – sin � cos �)<br />

P = 2R�<br />

R H =<br />

2� = �<br />

R H = D I/4<br />

R(� – sin � cos �)<br />

���<br />

2�<br />

2� > �<br />

� = � – �<br />

A = R 2 (� – sin � cos �)<br />

P = 2R�<br />

R H =<br />

R(� – sin � cos �)<br />

���<br />

2�<br />

FIGURE 6-2 Hydraulic radius for sha<strong>pe</strong>s of o<strong>pe</strong>n channels.<br />

B<br />

2R<br />

H<br />

H<br />

6.3<br />

if H > R<br />

A = R[2(H – R) + �R/2]<br />

R = 2(H – R) + �R<br />

R H =<br />

A = BH<br />

P = 2H + B<br />

R H = BH � 2H + B<br />

R[2(H – R) + �R/2]<br />

���<br />

2(H – R) + �R


6.4 CHAPTER SIX<br />

or<br />

fDU S = = (6-6)<br />

Many models for o<strong>pe</strong>n channel flows of water are based on the Chezy number and the<br />

Manning number. The Chezy number is inversely proportional to the square root of the<br />

friction factor:<br />

8g<br />

Ch = � (6-7)<br />

�� fD<br />

2 H<br />

� �<br />

L 8gRH<br />

or<br />

or<br />

2g<br />

Ch = � (6-8)<br />

�� fN<br />

The Manning number is a function of both the hydraulic radius and the friction factor:<br />

R H 1/6<br />

n = � (6-9)<br />

Ch<br />

1/6 RH ��2 �f N<br />

g �� n = �<br />

(6-10)<br />

Some ex<strong>pe</strong>rimental values for the Manning number “n” are shown in Table 6-1 as derived<br />

for water flows. These values are not correct for transportation of solids, particularly<br />

solids that introduce a new roughness factor that we will discuss. This table is presented<br />

as a reference for dirty water, very dilute mixtures, or decant water that are present in<br />

mining and tailings circuits but do not constitute real slurries.<br />

Green et al. (1978) summarized the research activities of the U.S. Army Corps of Engineers<br />

who derived the following relationship between the hydraulic radius and effective<br />

roughness of the channel (in USCS units):<br />

R H 0.1667<br />

n = (6-11)<br />

where<br />

RH = hydraulic radius in feet<br />

k = effective linear roughness in feet<br />

n = Manning number in ft –1/3 ���<br />

23.85 + 21.95 log(RH/ks) /sec<br />

The linear roughness ks (Table 6-2) is also used to compute the flow of water in o<strong>pe</strong>n<br />

channels. The Ministry of Transport (1969) recommends the following equation for flow<br />

of viscous liquids in o<strong>pe</strong>n channels:<br />

k 1.225(�/�)<br />

V = – �3�2�R� H�S� log� � + ���<br />

(6-12)<br />

14.8 RH RH �3�2�R� H�S�<br />

where<br />

� = absolute viscosity of fluid<br />

� = density of fluid<br />

S = slo<strong>pe</strong> or energy gradient<br />

g = acceleration due to gravity


SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

TABLE 6-1 Typical Values for the Manning Number “n” for Water Flows (Do Not<br />

Use for Slurries)<br />

Manning factor “n,” Manning factor “n,”<br />

Channel surface ft –1/3 s –1 m –1/3 s –1<br />

Glass, plastic, machined metal surface 0.011 0.016<br />

Smooth steel surface 0.008 0.012<br />

Sawn timber, joints uneven 0.014 0.021<br />

Corrugated metal 0.016 0.024<br />

Smooth concrete 0.0074 0.011<br />

Cement, plaster 0.011 0.016<br />

Concrete culvert (with connection) 0.009 0.013<br />

Glazed brick 0.009 0.013<br />

Concrete, timber forms, unfinished 0.014 0.0208<br />

Untreated gunite 0.015–0.017 0.022–0.0252<br />

Brickwork or dressed masonry 0.014 0.0208<br />

Rubble set in cement 0.017 0.0252<br />

Earth excavation, clean, no weeds 0.020 0.022<br />

Earth, some stones and weeds 0.025 0.037<br />

Natural stream bed, clean and straight 0.020 0.030<br />

Smooth rock cuts 0.024 0.035<br />

Channels not maintained 0.034–0.067 0.050–0.1<br />

Winding natural channels with pools and shoals 0.033–0.040 0.049–0.059<br />

Very weedy, winding, and overgrown natural rivers 0.075–0.150 0.111–0.223<br />

Clean alluvial channels with sediments 0.031 (d 75) 1/6 0.0561 (d 75) 1/6<br />

After Manning (1895) and Henderson (1990).<br />

Having read Chapters 1–3, the reader must have become aware that sizing pi<strong>pe</strong> involves<br />

a straightforward relationship between the flow rate, the cross-sectional area of the<br />

pi<strong>pe</strong>, and the required velocity. In the case of o<strong>pe</strong>n launders, particularly those of rectangular<br />

and U-sha<strong>pe</strong>, the main concern is to avoid spills. At certain bends, around certain<br />

obstacles, or at a sudden reduction of physical slo<strong>pe</strong>, the flow may slow down considerably<br />

and even spill out of the launder. For straight runs away from such bends and junctions<br />

of launders, o<strong>pe</strong>n conduits are designed to be one-third full. When pi<strong>pe</strong>s are used as<br />

o<strong>pe</strong>n launders, they are typically sized to be 50% full, but in the case of tenacious froth<br />

they may be sized to be 25% or 30% full (Figure 6-3).<br />

Designers often prefer to have a steep launder rather than to suffer a loss of time unblocking<br />

settled <strong>slurry</strong>. As the above equations indicate, the friction loss factor does de<strong>pe</strong>nd<br />

on the hydraulic radius, and larger launders tend to require less slo<strong>pe</strong> than small<br />

launders. Excessively steep launders tend to lose their liners through fast wear. Obtaining<br />

the correct slo<strong>pe</strong> without an excessive margin of safety is the correct approach to engineering.<br />

Example 6-1<br />

A <strong>slurry</strong> of unknown pro<strong>pe</strong>rties is flowing in a half full 457 mm (18 in) pi<strong>pe</strong> with wall<br />

thickness of 9.5 mm (0.375 in). The measured flow rate is 0.189 m 3 /s (3000 US gpm).<br />

The launder is inclined at a slo<strong>pe</strong> of 2%. Determine the friction factor and the Chezy and<br />

Manning numbers.<br />

6.5<br />

using d 75 size in feet using d 75 size in m


6.6 CHAPTER SIX<br />

TABLE 6-2 Recommended Values of Absolute Roughness in mm<br />

Classification<br />

Values of roughness ks, mm (*)<br />

Average<br />

effective<br />

roughness<br />

of launder<br />

(assumed clean and new unless otherwise stated) Good Normal Poor mm (**)<br />

Smooth<br />

Drawn nonferrous pi<strong>pe</strong>s of aluminum, brass, cop<strong>pe</strong>r, 0.003<br />

alkathene, glass, Pers<strong>pe</strong>x, HDPE<br />

Plastic pi<strong>pe</strong>, welded joints 0.146<br />

Plastic pi<strong>pe</strong>, flanged or coupled joints 2.292<br />

Fiberglass pi<strong>pe</strong> (FRP), flanged or coupled 2.292<br />

Metal<br />

Asbestos cement 0.015<br />

Spun bitumen-lined pi<strong>pe</strong> 0.03<br />

Spun concrete-lined pi<strong>pe</strong> 0.03<br />

Wrought iron pi<strong>pe</strong> 0.03 0.06 0.15<br />

Rusty wrought iron pi<strong>pe</strong> 0.15 0.6 3<br />

Uncoated steel pi<strong>pe</strong> 0.015 0.03 0.06 0.725<br />

Coated steel pi<strong>pe</strong> 0.03 0.06 0.15<br />

Rubber-lined steel pi<strong>pe</strong> 1.35<br />

Plastic-lined steel pi<strong>pe</strong> 0.350<br />

Galvanized iron 0.06 0.15 0.30 0.726<br />

Coated cast iron 0.06 0.15 0.30<br />

Tate relined pi<strong>pe</strong>s 0.15 0.3 0.6<br />

Riveted steel pi<strong>pe</strong>s (untuberculated)—(good = girth,<br />

riveted only; normal = full riveted, ta<strong>pe</strong>r, or<br />

cylinder joints; poor = full riveted, butt-strap joints)<br />

Riveted steel pi<strong>pe</strong>s (untuberculated)—plates < 6 mm 0.6 1.5 3<br />

Riveted steel pi<strong>pe</strong>s (untuberculated)—plates > 6 mm 1.5 3 6<br />

Concrete<br />

Class 4—Monolithic construction against oiled 0.06 0.15<br />

steel surface with no surface irregularities,<br />

smooth-surfaced precast pi<strong>pe</strong>lines with no shoulders<br />

or depressions at the joints<br />

Class 4a—Monolithic construction in units of 2 m or 0.15 0.3<br />

over with spigot and socket joints, or ogee joints<br />

pointed internally<br />

Class 3—Monolithic construction against steel, wet-mix, 0.3 0.6 1.5<br />

or spun pre-cast pi<strong>pe</strong>s, or with cement or asphalt<br />

coating<br />

Class 2—Monolithic construction against rough 0.6 1.5<br />

texture precast pi<strong>pe</strong> or cement gun surface (for<br />

very coarse textures, take � = size of aggregate in<br />

evidence)<br />

Class 1—Precast pi<strong>pe</strong>s with mortar squeeze at joints 3 6 3.63<br />

Smooth trowel led surface 0.3 0.6 1.5 1.35<br />

Lined concrete pi<strong>pe</strong> 0.73<br />

Unlined concrete pi<strong>pe</strong> 1.35


TABLE 6-2 Continued<br />

SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

Classification<br />

Values of roughness ks, mm (*)<br />

Average<br />

effective<br />

roughness<br />

of launder<br />

(assumed clean and new unless otherwise stated)<br />

Steel construction<br />

Good Normal Poor mm (**)<br />

Welded sections, unlined 1.35<br />

Rolled sections, unlined 0.73<br />

Rubber lined 1.35<br />

Plastic lined 0.73<br />

Plastic construction, free formed<br />

Clayware<br />

0.73<br />

Pitch fiber<br />

Pitch fiber<br />

0.003 0.03<br />

Glazed vitrified clay, very accurately lined joints 0.06<br />

Glazed vitrified clay in 1 m under 600 mm diameter 0.15 0.3<br />

Clayware, glazed vitrified clay in 1 m under 600 mm<br />

diameter<br />

0.3 0.6<br />

Clayware, glazed vitrified clay in 0.6 m under 300 mm<br />

diameter<br />

0.15 0.3<br />

Clayware, glazed vitrified clay in 0.6 m over 300 mm<br />

diameter<br />

0.3 0.6<br />

Clayware, butt jointed drain tile 0.6 1.5 3 1.35<br />

Clayware, glazed brickwork 0.6 1.5 3<br />

Clayware, brickwork, well pointed 1.5 3 6<br />

Clayware, old brickwork in need of pointing<br />

Mature foul sewers constructed of materials with<br />

roughness � when new, not exceeding those given<br />

for mature sewers<br />

15 30<br />

Slimed not more than 6 mm 0.6 1.5 3<br />

Lime incrustations, grease, or slime not more than<br />

25 mm thick, or even layer of fine sludge<br />

6 15 30<br />

Gritty solids, lying unevenly in inverts (higher figures<br />

relate to shoals of debris at Froude number of order<br />

of 0.3 to 0.5)<br />

Unlined rock tunnels<br />

60 150 300<br />

Granite and other homogeneous rocks 60 150 300<br />

Diagonally bedded slates (use values with design<br />

diameter)<br />

Earth channels<br />

300 600<br />

Straight uniform artificial channels 15 60 150<br />

Straight natural channels, free from shoals, boulders,<br />

and weeds<br />

150 300 600<br />

*Data from the Ministry of Technology, United Kingdom (1969), Hydraulics Research Pa<strong>pe</strong>r 4,<br />

Tables for the Hydraulic Design of Storm-drains, Sewers and Pi<strong>pe</strong>-Lines.<br />

**Data from Green (1978).<br />

6.7


6.8 CHAPTER SIX<br />

H = R<br />

Solution in Metric Units<br />

Q = 3,000 US gpm × 3.785 = 11,355 L/min = 0.1893 m 3 /s<br />

ID of pi<strong>pe</strong> = 438.15 mm<br />

Area for a half full pi<strong>pe</strong> = �/8 × 0.43815 2 = 0.0754 m 2<br />

Velocity = 0.1893/0.0754 = 2.51 m/s<br />

Wetted <strong>pe</strong>rimeter = � × 0.43815/2 = 0.688 m<br />

Hydraulic radius = A/P = 0.0754/0.688 = 0.1095<br />

Slo<strong>pe</strong> = 2%<br />

fN(2.51) 0.02 = = 2.93 fN 2<br />

fNV ��<br />

2 × 9.81 × 0.1095<br />

2<br />

�<br />

2gRH<br />

Fanning friction factor, f N = 0.0068<br />

The Darcy factor, f D = f N × 4 = 0.027<br />

The Chezy number, Ch = �� = �� = 53.71 m1/2 /s<br />

The Manning number, n = R H 1/6 /Ch = 0.1095 1/6 /53.714 = 0.0129 m –1/3 /s<br />

Solution in USCS Units<br />

H = Z/3<br />

Z<br />

2R<br />

2g<br />

� fN<br />

Q = 3000<br />

2 × 9.81<br />

�<br />

0.0068<br />

US gpm = 3000/7.4805 = 401.04 ft 3 /min<br />

ID of a pi<strong>pe</strong> = 17.25 in = 17.25/12 = 1.4375 ft<br />

Area of a half full pi<strong>pe</strong> A = (�/8)1.4375 2 = 0.8115 ft 2<br />

Velocity = P/A = 401.04/0.8115 = 494.21 ft/min (8.237 ft/s)<br />

P = �D/2 = � × 1.4375/2 = 2.258 ft<br />

H = Z/3<br />

FIGURE 6-3 Recommended degree of fill for o<strong>pe</strong>n channel <strong>slurry</strong> flows (does not apply to<br />

junction boxes).<br />

Hydraulic radius = A/P = 0.8115/2.258 = 0.359 ft<br />

Z<br />

B<br />

H


Slo<strong>pe</strong> = 2%<br />

0.02 = = = 2.935 fN The fanning friction factor, fN = 0.0068<br />

The Darcy factor, fD = fN × 4 = 0.027<br />

The Chezy number, Ch = = = 97.2 ft �� �� 1/2 fN × 8.237<br />

2g 2 × 32.2<br />

� � /s<br />

fN 0.0068<br />

1/6 1/6 –1/3 The Manning number, n = RH /Ch = 0.359 /97.2 = 0.0087 ft /s<br />

The reader should be careful when using the Manning roughness “n.” Contrary to<br />

common belief, it is not a nondimensional number and its value changes from SI units to<br />

USCS units by the ratio of conversion from feet to meters to the power of 1/3 or 0.673.<br />

2<br />

fNV ��<br />

2 × 32.2 × 0.359<br />

2<br />

�<br />

2gRH<br />

6-2 TRANSPORTATION OF SEDIMENTS<br />

IN AN OPEN CHANNEL<br />

Determining the Chezy number for slurries is a method of approaching the design of launders.<br />

Julian et al. (1921) measured an average Chezy number of 80 ft 1/2 /s for rectangular<br />

launders (with width = twice the depth) of minimum wetted <strong>pe</strong>rimeter for carrying slime<br />

overflow and average stamp-battery pulp in cyaniding gold and silver circuits.<br />

Classical theories of sus<strong>pe</strong>nded solids in o<strong>pe</strong>n channels are based on two-dimensional<br />

turbulent flow. Consider a two-dimensional turbulent flow with a velocity U in the horizontal<br />

x-direction and V in the vertical y-direction. Reynolds (1895) defined the shear<br />

stress parallel to x on a plane normal to y as<br />

where<br />

SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

� = –� (U�V�) average<br />

� = density of fluid and<br />

(U�V�) average = average of the fluctuations of the turbulent velocities<br />

(6.13)<br />

Boussinesq (1877) develo<strong>pe</strong>d an equation in the form of<br />

� = ��m (6-14)<br />

where<br />

��m = the eddy viscosity, analogous to the dynamic viscosity discussed in Chapter 2.<br />

�m = the coefficient of exchange of momentum between neighboring streams of the fluid,<br />

expressed in m2 /s or ft2 dU<br />

�<br />

dy<br />

/sec.<br />

Von Karman (1935) develo<strong>pe</strong>d the following equation:<br />

� = ��V�L mix<br />

dU<br />

�<br />

dy<br />

where<br />

� = correlation coefficient � 1.0 (see Section 6.2.3)<br />

V� = average of absolute values of fluctuations normal to the main flow<br />

L mix = mixing length<br />

6.9<br />

(6-15)


6.10 CHAPTER SIX<br />

By substituting Equation 6-15 into Equation 6-13 it is concluded that<br />

� m = �V�L mix<br />

A rate of transfer of mass of sus<strong>pe</strong>nded particles <strong>pe</strong>r unit area is defined as<br />

dm<br />

dC<br />

� = –�V�Lmix �<br />

dt<br />

dy<br />

(6-16)<br />

(6-17)<br />

But the values of �, V�, and Lmix are not necessarily equal in magnitude in Equations 6-14<br />

and Equation 6-17.<br />

C = concentration of sus<strong>pe</strong>nded solids<br />

O’Brien (1933) studied the sus<strong>pe</strong>nsion of sediments in an o<strong>pe</strong>n channel flow. He develo<strong>pe</strong>d<br />

a theory that the rate of transfer of solids upward must be in equilibrium with the<br />

downward exchange of momentum due to gravitational forces:<br />

dC<br />

VtCy = –�V�Lmix � (6-18a)<br />

dy<br />

VtCy = �s (6-18b)<br />

where<br />

Cy = volume concentration of solids at level y<br />

y = distance from the lower boundary<br />

�s = mass transfer coefficient for sediments, similar to �m but not necessarily equal to it.<br />

Solving Equation 6-18 yields<br />

loge = � y<br />

dC<br />

�<br />

dy<br />

C dC<br />

� � (6-19)<br />

Ca a dy<br />

where C a is the concentration of solids at an arbitrary reference plane of height “a.”<br />

If � s is constant over the depth, then � s(y) = constant. Equation 6-18 is then solved to<br />

give<br />

C<br />

� Ca<br />

= e –J (6-20)<br />

where J = (y – a)(V t/� s).<br />

The correct procedure consists of establishing a relationship between � s and the vertical<br />

coordinate y before solving Equation 6-18. In the absence of detailed information<br />

about the relationship between � s and � m, they are assumed to be equal so that<br />

�<br />

�<br />

�dU/dy<br />

� m = –�V�L mix = (6-21)<br />

Substituting Equation 6-21 into Equation 6-18 yields the concentration of distribution<br />

loge = –�Vt � y<br />

C dU/dy<br />

� � dy (6-22)<br />

Ca a �


In a uniform o<strong>pe</strong>n channel with a large ratio of width to depth, the shear stress is expressed<br />

as<br />

ym – y<br />

� = �w� � (6-23)<br />

where<br />

�w = shear stress at the wall<br />

ym = distance from the boundary to the liquid surface<br />

Substituting Equation 6-23 into Equation 6-21 yields<br />

loge = –�Vt � y<br />

C<br />

dU/dy<br />

� �� dy (6-24)<br />

Ca a (1 – y/ym)�w The velocity gradient is expressed in the form of the universal defect law as<br />

= log e� � (6-25)<br />

For a pi<strong>pe</strong>, K x = 0.4 and y is the distance from the internal wall at the bottom of a horizontal<br />

pi<strong>pe</strong>. Keulegan (1938) demonstrated that Equation 6-24 applied to o<strong>pe</strong>n channels.<br />

For a wide-o<strong>pe</strong>n channel, the value of y m, or depth of flow, is used:<br />

In Chapter 2, the friction velocity U f was defined as<br />

= log e� � (6-26)<br />

�w fN Uf = � = U � �� � ��2<br />

Substituting Equation 6-6 yields<br />

Uf = �(g�R� H�S�)� (6-27)<br />

where<br />

fN = the fanning friction factor<br />

U = the mean velocity of the flow<br />

S = slo<strong>pe</strong><br />

�w = �gRHS (6-28)<br />

Equation (6-28) is called the DuBoys equation (Wood, 1980). It clearly establishes that<br />

for a <strong>slurry</strong> to move at a density �, a minimum level of shear stress must be available:<br />

where<br />

SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

C<br />

� Ca<br />

U – Umax �<br />

�(��w/ ���)�<br />

U – Umax �<br />

�(��w/ ���)�<br />

V t<br />

� �KxU f<br />

log e = log e�� � � (6-29)<br />

C<br />

� Ca<br />

� ym<br />

1<br />

� Kx<br />

1<br />

� Kx<br />

h<br />

� ha<br />

= � � Z<br />

2y<br />

� DI<br />

y<br />

� ym<br />

ym – 1<br />

�<br />

y<br />

a<br />

�<br />

ym – a<br />

6.11<br />

(6-30)<br />

Vt Z = ��<br />

(6-31)<br />

�Kx�g�ym�S�


6.12 CHAPTER SIX<br />

h = � – 1 (6-32)<br />

y<br />

ym – a<br />

ha = � (6-33)<br />

a<br />

Equation 6-30 establishes that the relative concentration of solids de<strong>pe</strong>nds on their<br />

vertical position and on the factor Z, which is a function of the ratio of the terminal velocity<br />

of the particles Vt to the group �KxUf. It is therefore a measure of the intensity of turbulence.<br />

6-2-1 Measurements of the Concentration of Sediments<br />

y m<br />

Determining the magnitude of the plane “a” is the starting point to solve Equation 6-28<br />

and 6-29. Rouse (1937) suggested the height “a” be equal to the height of the roughness<br />

elements k s. He suggested using Equation 6-19, assuming an interval from y = 0 to y = k s,<br />

and by assuming that �(y) = �(k s) = constant. Rouse indicated that at y = 0 the solids concentration<br />

corresponding to the bed of sediments should be used.<br />

Richardson (1937) reported test results for a 305 mm × 305 mm × 1830 mm (1 ft × 1 ft<br />

× 6 ft) flume and indicated that in the boundary region the concentration of sediments was<br />

inversely proportional to the vertical coordinate y, but that in o<strong>pe</strong>n streams it conformed<br />

better with Equation 6-30.<br />

Vanoni (1946) conducted a series of tests in an 838 mm (33 in) wide by 18.29 m (60<br />

ft) long flume to validate Equation 6-29 (Figure 6-4). Average velocity was noticed to occur<br />

at 0.368 y m or the depth of the liquid. The velocity profile followed a logarithmic<br />

function of depth.<br />

The following results were obtained by Vanoni (1946).<br />

� The sediment concentration profile followed the pattern set by Equation 6-30. However,<br />

the exponent was smaller than the value of Z expressed by Equation 6-31 when the<br />

sediments became coarser.<br />

� A random turbulence was observed and slip between fluid and sediment was sus<strong>pe</strong>cted<br />

as the sediment accelerated. Thus, the assumption that mass and momentum transfer<br />

were equal was not satisfied, as the theory did not account for slip and random turbulence.<br />

� For fine materials, the coefficient of sediment mass transfer was smaller than the coefficient<br />

of momentum transfer.<br />

� For coarse materials, the coefficient of sediment mass transfer was larger than the coefficient<br />

of momentum transfer.<br />

� Sus<strong>pe</strong>nded load decreased the coefficient of mass transfer. The reduction was more<br />

important with fine solids than with coarse solids.<br />

� Sus<strong>pe</strong>nded load reduced the value of the Von Karman constant. Values of K x between<br />

0.314 and 0.342 were measured (by comparison with 0.4 for full pi<strong>pe</strong>s). The reduction<br />

of the Von Karman coefficient indicated a reduced level of mixing and a tendency by<br />

the sediments to suppress turbulence.<br />

� The sus<strong>pe</strong>nded load tended to reduce the resistance to flow. Sediment-laden water<br />

moved faster than clear water and the Manning roughness number decreased with the<br />

sediment load, as shown in Figure 6-5.


FIGURE 6-4 Velocity profile for the flow of a sand–water mixture in a rectangular o<strong>pe</strong>n<br />

channel. (From Vanoni, 1946, by <strong>pe</strong>rmission of ASCE.)<br />

6.13


6.14 CHAPTER SIX<br />

FIGURE 6-5 Variation of the equivalent roughness, Von Karman coefficient, and Z 1 with<br />

the weight concentration of the sand–water mixture. (From Vanoni, 1946, by <strong>pe</strong>rmission of<br />

ASCE.)


SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

� Sus<strong>pe</strong>nded load tended to cause a flow to become unevenly distributed. Unsymmetrical<br />

sediment distribution within the flow caused secondary circulation.<br />

� The velocity distribution near the center of the flume followed Von Karman universal<br />

defect law.<br />

Example 6-2<br />

Iron sand is flowing in a rectangular o<strong>pe</strong>n channel that is 300 mm wide × 120 mm high at<br />

a s<strong>pe</strong>ed of 3 m/s. Assume the Von Karman coefficient K = 0.33. The average particle size<br />

is 0.3 mm. Using Einstein’s approaches, it is assumed that the reference layer “a” is to be<br />

twice the particle size diameter. The flume is one-third full (i.e., y m = 40 mm). The <strong>slurry</strong><br />

weight concentration is 45%, the dynamic viscosity is 1 cP, and the s<strong>pe</strong>cific gravity of<br />

sand is 4.1. Calculate C/C a if the slo<strong>pe</strong> is 3% at depth with 2% increments of depth. Ignore<br />

any dunes. Assume � = 1.0.<br />

Solution in Metric Units<br />

The first approach is to determine the terminal velocity of the sand. The Particle Reynolds<br />

number is:<br />

d pV m�/� = (0.3 × 10 –3 × 3 × 1000/1 × 10 –3 ) = 900<br />

This is turbulent flow. By Newton’s law (Equation 3-13):<br />

V t = 1.74[g(� s – � L/� L)] 0.5 d p 0.5<br />

V t = 1.74[9.81 × 3.1 × 0.3 × 10 –3 ] 0.5<br />

Vt = 0.167 m/s<br />

Using Equation 6-31:<br />

Z = 0.167/[0.33 × (9.81 × 0.04 × 0.03) 0.5 ]<br />

Z = 4.664<br />

Applying Equation 6-29:<br />

C/Ca = (h/ha) 4.664<br />

As it will be explained later in this chapter, Einstein proposed that the value of “a” be<br />

equal to twice the grain diameter, or in this case 0.6 mm = twice the particle size of sand.<br />

Let us calculate the concentration of solids at 2% of the depth of the flume.<br />

2% of depth = 0.02 × 40 = 0.8 mm:<br />

y m<br />

h = � – 1 = h = (1/0.02) – 1 = 49<br />

y<br />

ym – a<br />

ha = �<br />

a<br />

ha = (40/0.6) – 1 = 65.67<br />

h<br />

� ha<br />

C<br />

� Ca<br />

= 49/65.67 = 0.735<br />

= 0.735 4.664 = 0.2378<br />

6.15


6.16 CHAPTER SIX<br />

4% of depth:<br />

C<br />

� Ca<br />

h<br />

� ha<br />

h = (1/0.04) – 1 = 24<br />

= 24/65.67 = 0.3655<br />

= 0.3655 4.664 = 9.15 × 10 –3<br />

So most of the solids will be in the bottom 4% of the launder.<br />

Example 6-3<br />

Iron sand (SG = 4.1) is flowing in a rectangular channel 600 mm wide × 120 mm high.<br />

The liquid level is 60 mm high. The slo<strong>pe</strong> is 1% and the liquid is flowing at 1.5 m/s. The<br />

particle average diameter is 0.5 mm. The s<strong>pe</strong>cific gravity of the solids is 4.1 and the dynamic<br />

viscosity of the mixture is 1.5 cP. Calculate C/C a at 4% intervals. Assume Von<br />

Karman K x = 0.33. Ignore any dunes. Assume � = 1.0.<br />

Solution<br />

The particle Reynolds number is :<br />

0.5 × 10 –3 × 1.5 m/s × 1000/1.5 × 10 –3 = 500<br />

This is transition flow. From Chapter 3, Allen’s law would apply:<br />

Vt = 0.20�g � 0.72<br />

�<br />

�L<br />

V t = 0.20(9.81 × 3.1) 0.72<br />

Vt = 1.116 mm/s<br />

Using Equation 6-30:<br />

Z = 1.116 × 10 –3 /[0.33 (9.81 × 60 × 10 –3 × 0.01) 0.5 ]<br />

Z = 3.3822 × 10 –3 /0.0767 = 0.044<br />

The magnitude of “a” is assumed to be twice the particle’s diameter:<br />

a = 2dp = 2 × 0.5 mm = 1 mm<br />

C<br />

� Ca<br />

� s/� L<br />

= (h/h a) 0.044<br />

d � 1.8<br />

� (�/�) 0.45<br />

(0.5 × 10 –3 ) 1.8<br />

���<br />

(1.5 × 10 –6 ) 0.45<br />

60 mm<br />

ha = �� = 59<br />

1 mm – 1<br />

Let us calculate the concentration at 2% depth:<br />

y = 0.02 × 60mm = 1.2 mm<br />

1<br />

h = � = 49<br />

0.02 – 1


4% depth:<br />

8% depth:<br />

12% depth:<br />

16% depth:<br />

20% depth:<br />

24% depth:<br />

SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

C<br />

� Ca<br />

C<br />

� Ca<br />

C<br />

� Ca<br />

C<br />

� Ca<br />

C<br />

� Ca<br />

= 49/59 = 0.83<br />

= 0.83 0.044 = 0.99<br />

y = 2.4 mm<br />

1<br />

h = � = 24<br />

0.04 – 1<br />

h<br />

= 24/59 = 0.4067<br />

= 0.4067 0.044 = 0.96<br />

y = 4.8 mm<br />

1<br />

h = � = 11.5<br />

0.08 – 1<br />

C<br />

� Ca<br />

h<br />

� ha<br />

C<br />

� Ca<br />

� ha<br />

= (11.5/59) 0.044 = 0.931<br />

y = 7.2 mm<br />

= (7.33/59) 0.044 = 0.912<br />

y = 9.6 mm<br />

= (5.25/59) 0.044 = 0.899<br />

y = 12 mm<br />

= (4/59) 0.044 = 0.888<br />

y = 14.4 mm<br />

= (3.17/59) 0.044 = 0.879<br />

6.17


6.18 CHAPTER SIX<br />

28% depth:<br />

y = 16.8 mm<br />

= 0.871<br />

32% depth:<br />

y = 19.2 mm<br />

= 0.864<br />

Examples 6-2 and 6-3 show the effect of slo<strong>pe</strong> on the distribution of solids. In the case<br />

of Example 6-2, the slo<strong>pe</strong> is higher at 3% and most of the solids move at the bottom of the<br />

channel. In Example 6-3, the slo<strong>pe</strong> is low at 1% and all the sand is mixed with the water.<br />

Graf (1971) reviewed the ex<strong>pe</strong>riments conducted by various researchers and a tendency<br />

develo<strong>pe</strong>d to measure an empirical Z1 as a substitute for Z. He summarized the work of<br />

Einstein and Chien (1955) who develo<strong>pe</strong>d an approximate relationship between Z and Z1: Z<br />

Z1 = �����<br />

(6-34)<br />

exp(–L2Z 2 2ZL<br />

/�) + �� � �(<br />

2� ��)�<br />

LZ�(2���)�<br />

exp(–x<br />

0<br />

2 C<br />

�<br />

Ca<br />

C<br />

�<br />

Ca<br />

/2)dx<br />

where<br />

x = log e y<br />

L = log e(1 + RK x)<br />

The best fit occurs when RK x = 0.3.<br />

Einstein (1950) called the flow layer right on top of the bed the “bed layer,” and indicated<br />

that it would be impossible to have sus<strong>pe</strong>nsion of the solids there. He measured a<br />

thickness of layer t = 2d p. The material within this layer was the source of the sus<strong>pe</strong>nded<br />

load and established the lower limit for C a. Einstein then proceeded to derive very complex<br />

equations that require numerical integration. It would be beyond the sco<strong>pe</strong> of this<br />

book to dwell on such equations.<br />

6-2-2 Mean Concentrations for Dilute Mixtures (C v < 0.1)<br />

Celik and Rodi (1991) published data suggesting that Einstein’s equations were overestimating<br />

the concentration at a = 2d �. For fairly dilute sus<strong>pe</strong>nsions with a volumetric concentration<br />

of less than 10%, which is common in a lot of applications, they proposed a<br />

more simplified approach, defining the sus<strong>pe</strong>nded sediment load q bs as<br />

where<br />

q bs = flow rate of sediment <strong>pe</strong>r unit width<br />

q b = total flow rate of mixture <strong>pe</strong>r unit width<br />

C = time averaged (mean) concentration<br />

C T = mean transport capacity concentration<br />

U = time averaged velocity in x-direction<br />

R H = hydraulic radius or mean depth of liquid<br />

qbs = qbCT = � RH �LUCdy (6-35)<br />

� a


SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

�a = depth of bed load layer<br />

y = vertical coordinate above bottom of channel<br />

Defining parameter Cm as the depth-averaged concentration, Celik and Rodi (1991)<br />

used their results from their previous pa<strong>pe</strong>r (1984) to conclude that Cm/CT � 1.13, and derived<br />

the following equation:<br />

CT = � ��<br />

(6-36)<br />

where � is a constant of proportionality (� 0.034 from tests on sand).<br />

The exercise consists of calculating C T. For this purpose, these two authors discussed<br />

the problem of the effective shear stress. Reviewing the work of other authors and their<br />

own, Celik and Rodi (1991) pointed out that there are certain important factors in the turbulent<br />

regime, such as:<br />

� When there are elements causing increased roughness, the flow separates, and only a<br />

part of the shear stress determined from the energy slo<strong>pe</strong> is effective in moving the<br />

particles in sus<strong>pe</strong>nsion. The work of gravity is then used partly to overcome friction at<br />

the bed.<br />

� The turbulent energy, which is produced in a turbulent regime, is related to the total<br />

shear stress, but the drag force is then not effective in maintaining the particles in sus<strong>pe</strong>nsion.<br />

� The turbulence energy of the up<strong>pe</strong>r layers needs to be transferred by convection and<br />

diffusion first to the region above the bed. Smaller quantities of energy are then available<br />

to sus<strong>pe</strong>nd the bed in the presence of large amounts of roughness.<br />

� The mean velocities and the wall shear stresses in separated regions are much smaller<br />

than in the areas where the flow is still attached.<br />

� The presence of separated regions in the flow and stagnant areas cause the particles to<br />

settle in dead water zones and it becomes very difficult to resus<strong>pe</strong>nd them.<br />

� The <strong>pe</strong>rmeability of a bed increases the resistance of the bed. (Zip<strong>pe</strong> and Graf, 1983).<br />

� For rivers, a typical value of the ratio of friction velocity to average velocity (U f /U) is<br />

0.05.<br />

Tests conducted by Van Rijn (1981) and by Apmann and Rumer (1967) indicate that<br />

the flow over an approximately flat bed of loose sand shows great similarity to the characteristics<br />

of flow over a rough surface.<br />

To take in to account all these effects in the turbulent regime, Celik and Rodi (1991)<br />

proposed an equation for the effective shear rate:<br />

�e = [1 – (ks/ym) � ]�w (6-37)<br />

where<br />

ks = the equivalent resistance parameter (in most cases the roughness height or absolute<br />

roughness)<br />

� = empirical constant (� = 0.06 in tests obtained by these authors)<br />

ym = average depth of liquid in flume<br />

In conclusion, Celik and Rodi (1984) established a simplified relationship between<br />

roughness and friction velocity for dilute mixtures as<br />

k s<br />

� ym<br />

�w Um �<br />

(�s – �L)gym Vt<br />

6.19<br />

= Er exp� –1 – K Um x �� (6-37)<br />

Uf


6.20 CHAPTER SIX<br />

where<br />

E r � 30<br />

K x = the Von Karman constant<br />

Substituting Equation 6-37 into Equation 6-36 and using the effective shear stress<br />

yields<br />

CT = �[1 – (ks/ym) � ] ��<br />

(6-39)<br />

Equation 6-39 is plotted in Figure 6-6.<br />

From test data, Celik and Rodi (1991) obtained a value where � = 0.034 and � = 0.06<br />

for flow over a flat bed of loose sand without large dunes or antidunes. The particle diameter<br />

was between 0.005 mm and 0.6 mm and volumetric concentration was limited to<br />

10%. On a logarithmic scale, the slo<strong>pe</strong> he obtained was 0.034 in the range of C T from<br />

k s<br />

�<br />

F = �1 – � � ym<br />

�� � w<br />

(�s – � L)gy m<br />

FIGURE 6-6 The volumetric capacity C T from Celik and Rodi (1991). (Reprinted by <strong>pe</strong>rmission<br />

of ASCE.)<br />

Um �<br />

Vt<br />

U m 2<br />

��<br />

(�s�� L – 1)gy m<br />

U m<br />

� Vt


0.00001 to 0.10. Nevertheless the authors pointed that the value of � is very empirical and<br />

strongly de<strong>pe</strong>ndent on the friction velocity U f. Unfortunately, the did not provide the correlation,<br />

and this leaves the engineer facing a situation of measuring such a value or iterating<br />

from case to case.<br />

Example 6-4<br />

Fine sand with an average particle diameter dp of 0.3 mm (0.000984 ft) and SG = 2.625 is<br />

transported at a volumetric concentration CV = 7.91%. The volume flow rate is 1200<br />

m3 /hr (42,378 ft3 /hr). The launder is 600 mm (1.97 ft) wide and the height of the liquid is<br />

200 mm. Determine the slo<strong>pe</strong> of the launder using the Celik–Rodi method. Assume Kx =<br />

0.33, � = 0.06, and � = 1.5 cP (or 3.13 × 10 –5 lbf-sec/ft2 ).<br />

Solution in SI Units<br />

Ct = = 0.07<br />

The average s<strong>pe</strong>ed is<br />

Um = = (1200/3600)/(0.2 × 0.6) = 2.79 m/s<br />

Assuming the roughness of the bed to be equal to twice the average particle diameter,<br />

ks = 0.6 mm<br />

From equation 6-38:<br />

= Er exp� –1 – K Cv �<br />

1.3<br />

Q<br />

�<br />

A<br />

ks Um � x �� ym<br />

Uf<br />

Using Equation 6-27:<br />

SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

2.79<br />

= 30 exp � –1 – 0.33 �<br />

Uf �<br />

ln(0.001) = 1 – � 0.9207<br />

�<br />

U<br />

U f = 0.112 m/s<br />

U f = �g�R� H�S�<br />

The hydraulic radius for this rectangular channel is<br />

(0.6 · 0.2)<br />

RH = �� = 0.12 m<br />

(0.6+0.2+0.2)<br />

U f = 0.1395 m/s = (9.81 × 0.12 × S) 1/2<br />

S = 0.0107 or 1.07%<br />

f<br />

6.21


6.22 CHAPTER SIX<br />

Solution in USCS Units<br />

The average s<strong>pe</strong>ed is<br />

Cv Ct = � = 0.07<br />

1.3<br />

Area of flow = 1.97 × 0.656 = 1.293 ft 2<br />

U m = = [(42378 ft 3 /hr)/3600]/1.293 = 9.1 ft/sec<br />

Assuming the roughness of the bed to be equal to twice the average particle diameter,<br />

If the depth is 8 in:<br />

From Equation 6-38:<br />

Using Equation 6-27:<br />

Q<br />

�<br />

A<br />

k s = 0.0236 in<br />

= 0.0236/8 = 0.003<br />

= Er exp� –1 – K Um x �<br />

9.1<br />

= 30 exp � –1 – 0.33 �<br />

Uf �<br />

ln(0.001) = 1 –<br />

U f = 0.38 ft/sec<br />

U f = �g�R� H�S�<br />

The hydraulic radius for this rectangular channel is<br />

6-2-3 Magnitude of ��<br />

k s<br />

� ym<br />

k s<br />

� ym<br />

R H = (1.987 · 0.65)/(1.987 + 0.65 + 0.65) = 0.391 ft<br />

U f = 0.38 ft/sec = (32.2 × 0.391 × S) 1/2<br />

S = 0.011 or 1.1%<br />

3.003<br />

� Uf<br />

Carstens (1952) demonstrated that � never exceeds unity (1.0). For fine particles, � � 1.0<br />

or � s � �. For coarse particles, � < 1.0 or � s < �.<br />

� Uf


SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

Brush et al. (1962), Matyukhin an Prokofyev (1966) and Majumdar and Carstens<br />

(1967) have confirmed that � = 1.0 or � s � � for fine particles. For coarse particles, � <<br />

1.0 or .� s < �.<br />

6-3 CRITICAL VELOCITY AND CRITICAL<br />

SHEAR STRESS<br />

6.23<br />

Graf (1971) established a relationship between the forces for the incipient movement of a<br />

set of loose, cohesionless solid particles and the angle of repose as<br />

FT tan � = �<br />

FN<br />

where<br />

FN = force normal to angle of repose<br />

FT = force tangential to angle of repose<br />

These two forces are the resultants of the lift and drag forces (discussed in Chapter 3) referred<br />

to in Figure 6-7:<br />

FN = W cos � – L<br />

FT = W cos � + D<br />

W cos � + D<br />

tan � = ��<br />

(6-40)<br />

W cos � – L<br />

The surface area resisting motion is expressed in terms of a sha<strong>pe</strong> factor � 1 and the<br />

particle diameter d �. The surface area associated with lift is expressed in terms of a sha<strong>pe</strong><br />

factor � 2 and the particle diameter d �:<br />

L = 0.5C L�U b 2 �2d � 2 (6-41)<br />

D = 0.5C D�U b 2 �1d � 2 (6-42)<br />

where U b is the bed velocity.<br />

The submerged weight of the particle is expressed as a sha<strong>pe</strong> factor and the diameter<br />

of the particle:<br />

W = � 3gd � 2 (�s – �) (6-43)<br />

Substituting Equations 6-40, 6-41, and 6-42 into 6-43 establishes the relationship between<br />

the critical velocity and the actual sha<strong>pe</strong> and density of the particle:<br />

U 2�3(tan � cos � – sin �)<br />

= ���<br />

(6-44)<br />

CD�1 + CL�2 tan �<br />

2 bc<br />

��<br />

(�s/�L – 1)<br />

where U bc is the the critical bed velocity to start the motion of the particles.<br />

Graf (1971) defined the right-hand part of Equation 6-44 as the sediment coefficient �.<br />

Fortier and Scobey (1926) conducted extensive ex<strong>pe</strong>riments on <strong>pe</strong>rmissible canal velocities<br />

to understand the erosion and transportation of sediments. Their results are presented<br />

in Table 6-3. Their main conclusions which are still valid today, were


6.24 CHAPTER SIX<br />

Drag<br />

Lift<br />

� The laws governing the transport of silt and detritus in o<strong>pe</strong>n channels are very distantly<br />

related to the laws governing scouring of the canal bed and are not directly applicable.<br />

� The material of seasoned canal beds consists of solids of different sha<strong>pe</strong>s and sizes.<br />

When the fines fill the interstices between the coarser solids, they form a dense and<br />

stable mass that is more resistant to the erosion of water.<br />

� The velocity required to scour the bedded canal is much higher than the velocity required<br />

to sus<strong>pe</strong>nd particles outside the bed.<br />

� Colloids tend to cement clay, sand, and gravel in such a way that the compound mixture<br />

resists erosion.<br />

� The grading of materials ranging from fine to coarse, coupled with the adhesion between<br />

colloids and these solids, makes it possible to o<strong>pe</strong>rate at high velocities without<br />

appreciable scouring effects.<br />

Neill (1967) derived the following equation to estimate the critical velocity for coarse<br />

particles in launders:<br />

U 2 bc<br />

��<br />

(�s/� – 1)gd P<br />

Weight<br />

= 2.50 � � –0.20<br />

(6-45)<br />

The term d P/y m is sometimes called “relative sand roughness.”<br />

In Chapters 3 and 4 we examined definitions of velocity during all <strong>pe</strong>riods of movement<br />

from settling to deposition, etc. Similarly, in o<strong>pe</strong>n channels the critical scour velocity is well<br />

above the sedimentation velocity (or equal to the terminal velocity in full pi<strong>pe</strong> flow). Sometimes<br />

engineers confuse these two terms, although they are quite different in magnitude.<br />

The critical shear stress is at the point of the incipient motion, or at which the motion<br />

starts, and is expressed as<br />

� cr<br />

��<br />

(�s – �)d �<br />

Velocity<br />

FIGURE 6-7 Lift and drag forces on sediments in an o<strong>pe</strong>n channel.<br />

= � (6-46)<br />

where � is the the sediment coefficient.<br />

Thomas (1979) argued that sand particles finer than 0.15 mm would be completely envelo<strong>pe</strong>d<br />

in the viscous sublayer so that the critical shear stress could be simplified to<br />

dP �<br />

ym


SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

6.25<br />

�cr = 1.21 �m[g�(�s/�L – 1)] 2/3 (6-47)<br />

With � = kinematic viscosity. Wilson (1980) indicated that this correlated well with the<br />

work of Ambrose (1953), who had indicated a change of resistance as the grains of sand<br />

became of the order of magnitude of the roughness of the flume.<br />

Substituting Equation 6-47 into the DuBoys equation (6-28), the critical slo<strong>pe</strong> to initiate<br />

motion of a bed of small sand particles in an o<strong>pe</strong>n channel is<br />

�w = �mgRHS = 1.21 �m[g�(�s/�L – 1)] 2/3<br />

TABLE 6-3 Permissible Canal Velocities (after Fortier et al., 1925)<br />

Velocity, after aging, of canals at a depth<br />

of 910 mm (3 ft) or less<br />

Water<br />

transporting<br />

noncolloidal<br />

Water silts, sands,<br />

Clear water, transporting gravel, or<br />

no detritus<br />

____________<br />

colloidal silts<br />

____________<br />

rock fragments<br />

_____________<br />

Original material excavated for the canal m/s ft/s m/s ft/s m/s ft/s<br />

Fine sand (noncolloidal) 0.45 1.5 0.84 2.5 0.45 1.5<br />

Sandy loam (noncolloidal) 0.54 1.75 0.84 2.5 0.61 2.0<br />

Silt loam (noncolloidal) 0.61 2.0 0.91 3.0 0.61 2.0<br />

Alluvial silts when noncolloidal 0.61 2.0 1.07 3.5 0.61 2.0<br />

Ordinary firm loam 0.84 2.5 1.07 3.5 0.69 2.25<br />

Volcanic ash 0.84 2.5 1.07 3.5 0.61 2.00<br />

Fine gravel 0.84 2.5 1.52 5.0 1.15 3.75<br />

Stiff clay (very colloidal) 1.15 3.75 1.52 5.0 0.91 3.0<br />

Graded loam to cobbles, when noncolloidal 1.15 3.75 1.52 5.0 1.52 5.0<br />

Alluvial silts when colloidal 1.15 3.75 1.52 5.0 0.91 3.0<br />

Graded, silt to cobbles, when colloidal 1.22 4.0 1.67 5.5 1.52 5.0<br />

Coarse gravel (noncolloidal) 1.22 4.0 1.83 6.0 1.98 6.5<br />

Cobbles and shingles 1.5 5.0 1.67 5.5 1.98 6.5<br />

Shales and hard pans 1.83 6.0 1.83 6.0 1.5 5.0<br />

TABLE 6-4 Effects of Sediment Load on Equivalent Roughness<br />

(after Vanoni, 1946)<br />

Ratio of equivalent roughness<br />

Average sediment load in grams/liter to size of bottom sand<br />

0 0.328<br />

0.17 0.282<br />

3.21 0.190<br />

7.36 0.110<br />

16.2 0.072


6.26 CHAPTER SIX<br />

The critical slo<strong>pe</strong> to start motion is<br />

1.21[�(�s/�L – 1)]<br />

Scrit = (6-48)<br />

2/3<br />

���<br />

RHg1/3 Equation (6-47) confirms the correlation between the average depth of the <strong>slurry</strong>, the<br />

weight of solids, the viscosity and density of the mixture, and the minimum slo<strong>pe</strong>.<br />

Example 6-5<br />

A <strong>slurry</strong> mixture of fine particles and water has a s<strong>pe</strong>cific gravity of solids 4.1 and is<br />

flowing in a partially filled rectangular launder 600 mm wide. Determine the critical slo<strong>pe</strong><br />

if the launder is to flow one-third full. The density of the mixture is 1250 kg/m3 (2.42<br />

slugs/ft3 ) and the dynamic viscosity is 20 mPa · s (4.17 × 10 –4 lbf-sec/ft2 ).<br />

Solution in Metric Units<br />

The kinematic viscosity of the mixture is 20 × 10 –3 /1250 = 0.000016 m2 /s. If the launder<br />

is one-third full, the hydraulic radius is<br />

From Equation 6-47:<br />

0.6 × 0.2<br />

RH = � = 0.12 m<br />

0.6 + 0.4<br />

1.21[16 × 10<br />

Scrit = = 0.0063 or 0.63%<br />

–6 (4.1 – 1)] 2/3<br />

���<br />

0.12 × 9.811/3 1.21[�(�s/�L – 1)] 2/3<br />

��<br />

RHg1/3 is the minimum slo<strong>pe</strong> to start motion of the <strong>slurry</strong> in these conditions.<br />

Solution in USCS Units<br />

The width of the launder is 1.97 ft, the height of the liquid would be 0.66 ft, and the hydraulic<br />

radius is<br />

The kinematic viscosity of the mixture is<br />

From Equation 6-47<br />

4.17 × 10 –4 lbf-sec/ft 2<br />

���<br />

2.42 slugs/ft 3<br />

1.97 × 0.66<br />

RH = �� = 0.395 ft<br />

1.97 + 0.66<br />

= 1.723 × 10 –4 lbf-sec-ft/slugs<br />

1.21[1.723 × 10<br />

Scrit = = = 0.0063 or 0.63%<br />

–4 (4.1 – 1)] 2/3<br />

����<br />

0.395 × 32.21/3 1.21[�(�s/�L – 1)] 2/3<br />

���<br />

RHg1/3 is the minimum slo<strong>pe</strong> to start motion of the <strong>slurry</strong> in these conditions.<br />

This analysis was extended by Wilson (1980) for sand flowing in partially filled pi<strong>pe</strong>s.<br />

Wilson defined three regimes for sand flowing in o<strong>pe</strong>n launders with particle diameters in<br />

the range of 0.02 mm to 4 mm (mesh 625–5):


1. Homogeneous flow occurs at a slo<strong>pe</strong> S < 0.0006/D I<br />

2. Optimum slo<strong>pe</strong> for heterogeneous flow S = 0.10(d p/1000) 1.5 D I –0.7<br />

3. Fully blocked pi<strong>pe</strong> occurs at S > 0.42<br />

For rocks with solid s<strong>pe</strong>cific gravity between 1.5 and 6.0, Wilson (1980) extended the<br />

analysis and stated, for homogeneous flow:<br />

or heterogeneous flow:<br />

for blocked flow:<br />

SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

S1 � × 10 –3 0.6 [�(�s/�L – 1)]<br />

(6-49)<br />

0.6<br />

���<br />

DI 1650<br />

31.6 × 10<br />

S1 < S2 � (6-50)<br />

–3 1.5 d p [(�s/�L – 1)] 1.2<br />

����<br />

0.42[(�s/�L – 1)]<br />

S3 � (6-51)<br />

0.35<br />

��<br />

Example 6-6<br />

A <strong>slurry</strong> mixture of fine particles and water of d85 = 0.08 mm with a density of solids<br />

4100 kg/m3 is flowing in a partially filled circular pi<strong>pe</strong> with an inner diameter of 438 mm<br />

(17.25 in). The pi<strong>pe</strong> inner diameter is 438 mm (17.25 in). Determine the minimum slo<strong>pe</strong><br />

for flow as a heterogeneous mixture and the slo<strong>pe</strong> for a blocked pi<strong>pe</strong>.<br />

Solution from Equation 6-50<br />

For heterogeneous flow:<br />

S 2 �<br />

S 2 =<br />

S 2 = 0.086 or 8.6%<br />

The slo<strong>pe</strong> for a blocked pi<strong>pe</strong> is determined from Equation 6-51 as 52.4%.<br />

6-4 DEPOSITION VELOCITY<br />

D I 0.7 1.65 1.2<br />

1.65 0.35<br />

31623d p 1.5 [(�s/� L – 1)] 1.2<br />

���<br />

D I 0.7 1.65 1.2<br />

316230.00008 1.5 [(4.1- 1)] 1.2<br />

���<br />

0.438 0.7 1.65 1.2<br />

6.27<br />

The terrains over which tailing flumes are built do not always have an appropriate slo<strong>pe</strong>.<br />

If the flow o<strong>pe</strong>rates in a subcritical flow regime, the engineer must calculate a realistic estimation<br />

of the deposition velocity. Dominguez et al. (1996) published an equation based<br />

on ex<strong>pe</strong>rimental data measured at Codelco and the Chilean Research Center of Mining<br />

and Metallurgy. For cases where the viscosity effects are negligible


6.28 CHAPTER SIX<br />

� � 0.158<br />

VD = 1.833� � 1/2<br />

8gRH (�S – �m) d85 ��<br />

(6-52)<br />

However, in cases where the dynamic viscosity of the carrier liquid is instrumental,<br />

such as with alkaline water, Dominguez et al. (1996) derived the following equation:<br />

VD = 1.833� � 1/2<br />

8gRH (�S – �m) ��<br />

where<br />

J = R H(gR H) 1/2 /� m<br />

� m = the absolute viscosity of the mixture<br />

� � 0.158 d85 1.2 (3,100/J) (6-53)<br />

A comparison between the deposition velocity as calculated by Equation 6-52 and ex<strong>pe</strong>rimental<br />

data is presented in Figure 6-8.<br />

Example 6-7<br />

A <strong>slurry</strong> mixture of coarse particles of d 85 = 12 mm with C w = 40%, density of solids 4100<br />

kg/m 3 , and a s<strong>pe</strong>cific gravity 4.1 is flowing in a circular pi<strong>pe</strong> with an inner diameter of 438<br />

mm (17.25 in). Assuming that the pi<strong>pe</strong> is to be half full, determine the deposition velocity.<br />

Solution<br />

Using Equation 6-51:<br />

�m<br />

�m<br />

� RH<br />

� RH<br />

R H = D/4 = 0.438/4 = 0.1105 m (4.35 in)<br />

� m = 1000/[1 – 0.4(4100 – 1000)/4100] = 1433kg/m 3<br />

FIGURE 6-8 Correlation between calculated and ex<strong>pe</strong>rimental measurements of the deposition<br />

velocity of coarse slurries. (From Dominguez et al., 1996, reprinted by <strong>pe</strong>rmission of BHR<br />

Group.)


SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

V D = 1.833[8 × 9.81 × 0.1105 (4100 – 1433)/1433] 0.5 (0.012/0.1105) 0.158<br />

VD = 1.29 × 4.0174 = 5.183 m/s<br />

Green et al. (1978) and the ASCE and WPCF (1977) proposed that Camp’s equation<br />

for the self-cleaning s<strong>pe</strong>ed of sewers be used in the design of <strong>slurry</strong> launders:<br />

Vsc = � dpg� �� 1/2<br />

8Ke �s – �m � �<br />

fD �m<br />

(6-54)<br />

where<br />

Ke = an ex<strong>pe</strong>rimental constant (the ASCE recommends a constant of 0.06 for grit chambers,<br />

whereas Green recommends a constant of 0.8 for sewers)<br />

Vsc is expressed in m/s (ft/s)<br />

d� is expressed in m (ft)<br />

g = 9.81 m2 /s (32 ft/s)<br />

To use this equation, one must first determine the Darcy friction factor and then solve<br />

by iteration. Because the average s<strong>pe</strong>ed is a logarithmic distribution of depth, it would be<br />

wise to design launders with a mean s<strong>pe</strong>ed equal to twice the self-cleaning s<strong>pe</strong>ed.<br />

Example 6-8<br />

Calculate the self-cleaning s<strong>pe</strong>ed of Example 6-1, assuming a sewer flow where K e = 0.8,<br />

solids at a SG of 3.1, C w = 20%, and d p = 2 mm.<br />

Solution in Metric Units<br />

In Example 6-1, the Darcy factor was calculated to equal f D = f N × 4 = 0.027.<br />

� m = 1000/[1 – 0.2(3100 – 1000)/3100] = 1157 kg/m 3<br />

V sc = [8 × 0.8 × 2 × 10 –3 × 9.81(3100 – 1157)/(1157 × 0.027)] 1/2<br />

V sc = 2.79 m/s<br />

Solution in USCS Units<br />

In Example 6-1, the Darcy factor was calculated to equal fD = fN × 4 = 0.027.<br />

dp = 6.56 × 10 –3 ft<br />

Sm = 1/[1 – 0.2(3.1 – 1)/3.1] = 1.157<br />

Vsc = [8 × 0.8 × 6.56 × 10 –3 × 32.2(3.1 – 1.157)/(1.157 × 0.027)] 0.5<br />

V sc = 9.2 ft/s<br />

6-5 FLOW RESISTANCE AND FRICTION<br />

FACTOR FOR HETEROGENEOUS<br />

SLURRY FLOWS<br />

6.29<br />

Equations 6-6 to 6-10 established the parameter for friction loss of single-phase Newtonian<br />

flows in o<strong>pe</strong>n channels. Equation 2-25 established the correlation between friction<br />

factor and friction velocity. Two approaches are often used by designers of launders; one<br />

is based on the friction factor and the other on the velocity.


6.30 CHAPTER SIX<br />

6-5-1 Flow Resistances in Terms of Friction Velocity<br />

Liu (1957) and Acaroglu (1968) indicated that the ratio of the friction velocity to the settling<br />

s<strong>pe</strong>ed of a particle were implicitly related and proposed the following function:<br />

Uf d�Uf � = fct��� (6.55)<br />

Vt �<br />

Graf (1971) proposed including the Froude number to reflect the degree of turbulence:<br />

Uf d�Uf Uav � = fct� � ’ �<br />

� �(g�ym�)� �<br />

Vt<br />

where y m is the average depth of the fluid. This is confirmed in studies by Garde and Dattari<br />

(1963) and Bogardi (1965).<br />

The presence of sand dunes at the bottom of a channel with a typical wavelength � is a<br />

function of the Froude number (Kennedy, 1963). As shown in Figure 6-9, antidunes are<br />

formed in the regime of critical flow (0.8 < Fr < 1.5).<br />

The presence of dunes increases the effective wall shear stress in the form of profile<br />

drag. Graf (1971) proposed to establish a hydraulic radius R H� due to the grain roughness<br />

and a separate value R H�� based on the bed forms so that:<br />

� w = �gS(R H� + R H��)<br />

where S is the physical slo<strong>pe</strong>. He defined a friction velocity U f as a combination of the<br />

component due to grain roughness U f� and due to the bed form as U f��:<br />

Froude numbe number Fr<br />

F<br />

2.8<br />

2.4<br />

2.0<br />

1.6<br />

1.2<br />

0.8<br />

0.4<br />

U f 2 = Uf� 2 + U f�� 2<br />

antidunes<br />

dunes<br />

0.0<br />

0 2 4 6 8 10 12 14 16 18<br />

� = wavelength<br />

2 * * D/<br />

FIGURE 6-9 Wavelength of dunes and dunes versus the Froude number in o<strong>pe</strong>n channel<br />

flows. (After Kennedy, 1963.)


6-5-2 Friction Factors<br />

SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

6.31<br />

The previous paragraph demonstrates that the presence of the bed forms (dunes and antidunes)<br />

tends to increase the friction velocity and therefore the friction factors.<br />

6-5-2-1 Effect of Roughness<br />

There is a dearth of information on the effect of roughness factors on <strong>slurry</strong> flow in closed<br />

or o<strong>pe</strong>n conduits. The Ministry of Technology of the United Kingdom (1969) recommended<br />

modifying the Colebrook equation by using the hydraulic radius for the single-phase fluids.<br />

Green et al. (1978) concurred with this approach and proposed that the Darcy friction<br />

factor be obtained from the following equation by replacing the diameter with four times<br />

the hydraulic radius and using the Reynolds number based on the hydraulic radius:<br />

1<br />

ks 2.51<br />

� = –2 log10� � + �� (6-56)<br />

�fD�<br />

14.8RH Re �fD�<br />

where<br />

ks = the linear roughness (measured in the same units as the hydraulic radius, e.g., meters)<br />

Re = 4RHV�/�, the Reynolds number expressed in terms of the hydraulic radius<br />

This definition of the linear roughness is difficult to calculate. In a fast flow, the<br />

roughness of the pi<strong>pe</strong> or channel wall may be used. Attempts have been made to define a<br />

(Nikuradse) sand roughness for closed conduits, such as the ratio of the particle diameter<br />

to the inner diameter of the conduit (dp/DI) but very little has been published for o<strong>pe</strong>n<br />

channels. The problem is far from simple, and the roughness is often taken as twice the<br />

grain diameter. The presence of dunes at low Froude number tends to complicate the picture<br />

by introducing another parameter for roughness.<br />

Example 6-9<br />

Considering Example 6-1, assume that the roughness is 0.0045 mm. Reiterate the friction<br />

factor using Equation 6-56. Assume �m = 1350 kg/m3 and � = 2.8 cP.<br />

fD = 0.027<br />

RH = 0.1095 m<br />

V = 2.51 m/s<br />

Re = �V × 4RH/� = 1350 × 2.5 × 4 × 0.1095/2.8 × 10 –3<br />

Re = 131,987<br />

Iteration 1<br />

0.027 –0.5 = –2log10[0.0045/(14.8 × 0.1095) + (2.51/(131,987 · 0.0270.5 )]<br />

6.085 � 5.077<br />

Iteration 2<br />

Assume fD = 0.038, then<br />

0.038 –0.5 = –2log10(2.777 × 10 –3 + 3.707 × 10 –6 )<br />

5.13 � 5.11<br />

Therefore the Darcy factor is iterated to 0.038.<br />

6-5-2-2 Effect of Particle Concentration on Slurry Viscosity<br />

Green et al. (1978) proposed to incorporate the effect of particle concentration in the form<br />

of increased effective viscosity by using the Einstein–Thomas equation (Equation 1-9) to


6.32 CHAPTER SIX<br />

define an effective viscosity. This effective viscosity is then used to compute the Darcy<br />

factor by using the hydraulic radius in the Colebrook equation. This approach is, however,<br />

essentially limited to Newtonian slurries in pseudohomogeneous flow well above the<br />

deposition velocity. This approach has been covered by Equation 6-53. It does not take in<br />

account any dunes or partial deposition at the bottom of the bed . It is, however, a useful<br />

and straightforward approach for flows at su<strong>pe</strong>rcritical Froude number.<br />

6-5-2-3 Effects of Particle Sizes on the Chezy Coefficient<br />

Richardson et al. (1967) derived the following equations. For a plane with little or no sediment<br />

transportation:<br />

= 5.9 log� � + 5.44 (6-57)<br />

For a plane with appreciable sediment transport:<br />

= 7.4 log � � (6-58)<br />

For ripples (in English units):<br />

= �7.66 – � log Ch<br />

ym � �<br />

�g� d85<br />

Ch<br />

ym � �<br />

�g�<br />

d85<br />

Ch 0.3 ym 0.13<br />

� � � � � + � + 11 (6-59)<br />

�g� Uf d85 Uf<br />

For dunes and antidunes:<br />

Ch<br />

�<br />

�g�<br />

ym �<br />

d85<br />

�RHS �<br />

RHS<br />

where �RHS is the increase of RHS due to the form roughness.<br />

= 7.4 log � ��� 1� –����� (6-60)<br />

Example 6-10<br />

A launder is designed to transport appreciable coarse sediments over a plane with d85 =<br />

4 mm. The height of the <strong>slurry</strong> must be limited to 150 mm. Using Equation 6-60, determine<br />

the required slo<strong>pe</strong> if the flow rate is 850 m3 /hr and the width of the channel is<br />

450 mm.<br />

Solution<br />

Ch/�g� = 7.4 log(0.15/0.004) = 11.65<br />

Ch = 36.5 m1/2 /s<br />

Area of flow = 0.15 × 0.45 = 0.0675 m2 Q = 850 m3 /hr = 0.236 m3 /s<br />

V = 3.498 m/s<br />

fD = 8g/Ch2 = 8 × 9.81/36.52 = 0.0589<br />

Slo<strong>pe</strong> = fDV 2 /8gRH RH = A/P = (0.15 × 0.450)/(0.450 + 0.15 + 0.15)<br />

RH = 0.793 m<br />

S = 0.0589 × 3.4982 /[8 × 9.81 × 0.793] = 0.0116 or 1.16%


SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

6-5-2-4 Effect of Bed Form on the Friction<br />

Graf (1971) discussed the importance of Equation 6-52 and proposed to write the total<br />

friction factor for flow in a channel in the presence of dunes or bed forms. He suggested<br />

the following equation for the overall friction factor:<br />

fD = f D� + f D�� (6-61)<br />

where<br />

f D� = Darcy friction factor for the channel without bed forms<br />

f D�� = Darcy friction factor due to bed forms<br />

In concordance with Silberman (1963) and Vanoni and Hwang (1967), Graf (1971) indicated<br />

that f D� can be estimated from conventional pi<strong>pe</strong> equations by substituting the diameter<br />

of the pi<strong>pe</strong> with four times the hydraulic radius. This has already been presented in<br />

Equation 6-54. A similar approach was develo<strong>pe</strong>d by Lovera and Kennedy (1969).<br />

From lab tests, Vanoni and Hwang (1967) derived the following equation for f D��:<br />

= 3.5 log � – 2.3 (6-62)<br />

e�Hav<br />

where<br />

�Hav = mean height of the bed form<br />

e = A/Ab = (where A = total area and Ab = the horizontal projection of the lee face of<br />

the bed forms)<br />

When the magnitude of “e” cannot be determined, Equation 6-62 can be written in<br />

terms of the wavelength of the bed form as:<br />

1<br />

1<br />

�<br />

�f��� D�<br />

�<br />

�f��� D�<br />

= 3.3 log � – 2.3 (6-63)<br />

(�Hav) 2<br />

where � is the length of the dune. In this section, the importance of dunes was well emphasized.<br />

The designer of a <strong>slurry</strong> flume should avoid these troublesome regimes by designing<br />

for su<strong>pe</strong>rcritical flows wherever the topography allows it.<br />

6-5-3 The Graf–Acaroglu Relation<br />

Starting from basic principles of lift and drag forces and buoyancy and weight on a solid<br />

particle, Acaroglu (1968), Graf and Acaroglu (1968) proceeded to develop a methodology<br />

that applies equally well to both closed conduits and o<strong>pe</strong>n channels. They considered<br />

that the drag force was a function of the Reynolds number based on the bed velocity Ub and the sha<strong>pe</strong> factor �1: 2 2<br />

D = 0.5CD�LUb�1d �<br />

However, they chose a different sha<strong>pe</strong> factor �D for the drag coefficient:<br />

Ubdp CD = f1� � , �D� (6-64)<br />

�<br />

The bed velocity for the solids-water mixture is expressed as<br />

Uf ks �<br />

�<br />

R H<br />

�R H<br />

y<br />

6.33<br />

Ub = Uf f2� , CV, �� (6-65)<br />

ks


6.34 CHAPTER SIX<br />

where<br />

ks = the Nikuradse’s equivalent sand roughness (d�/DI) CV = the volumetric concentration of solids<br />

The shear friction velocity is expressed as<br />

Uf = � = �R� �� H�S�g� (6-66)<br />

�<br />

By assuming that the absolute roughness of the bed is equal to the particle diameter, Acaroglu<br />

and Graff proceeded to define the shear intensity parameter as<br />

(�s – �L)d� �A = ��<br />

(6-67)<br />

At a critical value of the shear intensity parameter � Acr, the shear stress is equal to the<br />

critical shear stress previously discussed. When � A > � Acr or � w < � cr, no movement of<br />

sediments occurs. When � A < � Acr or � w > � cr, movement of sediments occurs.<br />

The power consumed with friction or head losses in the o<strong>pe</strong>n channel is expressed in<br />

terms of the energy slo<strong>pe</strong> (head loss <strong>pe</strong>r unit length) and a nondimensional transport parameter<br />

is derived as<br />

�A = ��<br />

(6-68)<br />

By examining data from various authors and by regression analysis, Graf and Acaroglu<br />

extrapolated the following relationship:<br />

�A = 10.39 (�A) –2.52 (6-69a)<br />

or<br />

C VU avR H<br />

��<br />

�(��s/��� L�–� 1�)g�d� � 3 �<br />

�LSR H<br />

CVUavRH 3 �(��s/��� L�–� 1�)g�d� p� = 10.39� � –2.52<br />

(�s – �L)d� ��<br />

�LSR H<br />

(6-69b)<br />

Equation 6-66 was obtained for finely graded sand with a particle diameter between<br />

0.091 mm and 2.70 mm (0.0036 – 0.1063 in) and was studied in rivers and o<strong>pe</strong>n channel<br />

flumes. This equation applied to both closed conduits and o<strong>pe</strong>n channels (Figure 6-10).<br />

Graf (1971) pointed out that this equation was based on extensive data that was often difficult<br />

to analyze. For some unknown reasons, there has not been much research since<br />

1970 to refine the Graf–Acaroglu equation. This is probably due to the fact that practically<br />

all research on <strong>slurry</strong> flows tends to limit itself to full pi<strong>pe</strong>s.<br />

Example 6-11<br />

A mine is located at high altitude. The tailings need to be transported by gravity over a<br />

long distance. The estimated particle size is 6 mm. The volumetric concentration is set at<br />

25%. The s<strong>pe</strong>cific gravity of the solids is 3.1. The carrier liquid is water. The channel is<br />

rectangular in sha<strong>pe</strong> with a width of 750 mm. Assuming an average s<strong>pe</strong>ed of 2 m/s, determine<br />

the minimum slo<strong>pe</strong> for a flow rate of 1100 m 3 /hr using the Graf–Acaroglu method.<br />

Solution in Metric Units<br />

Q = 1100 m3 /hr or (1100/3600) = 0.3055 m3 /s<br />

For a s<strong>pe</strong>ed of 2 m/s, the height of the liquid would be:<br />

0.3055/(2 × 0.75) = 0.204 m<br />

� w


A<br />

10.0<br />

1.00<br />

0.10<br />

S<br />

L<br />

100.0<br />

SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

S R<br />

0.01<br />

The hydraulic radius is<br />

RH = (0.75 × 0.204)/(0.75 + 2 × 0.204) = 0.132 m<br />

From Equation 6-68:<br />

�A = = = 0.188<br />

From Equation 6-69a:<br />

–2.52 �A = 10.39 � A or �A = 4.911<br />

From Equation 6-67:<br />

4.911 × 0.132 = 3.1 × 6 × 10 –3s –1<br />

0.25 × 2 × 0.132 0.0662<br />

��� �<br />

�[( �3�.1� –� 1�)�9�.8�1� ×� 6� ×� 1�0� 0.3515<br />

–3 �]�<br />

S = 0.0287 or 2.87%.<br />

The Graf-Acaroglu method is very useful to determine the bed depth of a full pi<strong>pe</strong><br />

with saltation. Equation 4-43 of Chapter 4 uses this method.<br />

6-5-4 Slip of Coarse Materials<br />

L<br />

H<br />

d<br />

p<br />

0.1 1.0<br />

6.35<br />

Kuhn (1980) conducted velocity measurements on transportation of coarse coal in<br />

tra<strong>pe</strong>zoidal flumes under controlled laboratory conditions. He reported slip between the<br />

coarse solids and liquid. At the inclination of 2.5°, the s<strong>pe</strong>ed of the coarse material was<br />

of 10% slower than the liquid s<strong>pe</strong>ed of 4.5 m/s, but it decreased gradually to a slip of<br />

8.5% of the s<strong>pe</strong>ed of 6 m/s at an inclination of 6°. Such a degree of slip is reminiscent<br />

10<br />

A<br />

100<br />

C U R<br />

V<br />

S L<br />

FIGURE 6-10 The Graf–Acaroglu relationship for flow of sand in o<strong>pe</strong>n and closed channels,<br />

assuming a particle roughness equal to the grain size. Adapted from Graf and Acaroglu, 1968.<br />

H<br />

1000<br />

g d<br />

3<br />

p


6.36 CHAPTER SIX<br />

of the two-layer models of heterogeneous flow extensively discussed in Chapter 4 for<br />

full pi<strong>pe</strong> flows.<br />

6-5-5 Comparison between Different Models<br />

Blench et al. (1980) conducted a comparison between the different equations to calculate<br />

the sediment discharge versus the water discharge in o<strong>pe</strong>n channels and plotted them as in<br />

Figure 6-11. It is obvious that different equations yield different results. The sediments<br />

are transmitted in different patterns. When the flow is tranquil (Froude number Fr < 1),<br />

two kinds of sand waves may develop: dunes and ripples. They are similar in sha<strong>pe</strong>, with<br />

an upstream surface and a gentle and gradually varying slo<strong>pe</strong>, finishing with an abrupt<br />

downstream slo<strong>pe</strong> (Figure 6-12). Although similar in sha<strong>pe</strong>, ripples are inde<strong>pe</strong>ndent of<br />

the magnitude of flow, whereas dunes are strongly de<strong>pe</strong>ndent on flow.<br />

At Froude number larger than unity (sometimes referred to as su<strong>pe</strong>rcritical flows), the<br />

flow suppresses the formation of bed forms. Anti-dunes, which are more symmetrical than<br />

ripples, form. They move in the same direction as the flow or even opposite to the flow.<br />

FIGURE 6-11 Comparison between the different models for transport of sediments in o<strong>pe</strong>n<br />

channels. (From ASCE, 1975. Reprinted by <strong>pe</strong>rmission of ASCE.)


y m<br />

SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

Flow direction<br />

wavelength<br />

liquid level<br />

FIGURE 6-12 Sand dunes at low Froude number (Fr < 1).<br />

Anti-dunes accentuate the deformation of the free surface (Figure 6-13). They do not occur<br />

in closed conduits and their motion is at a lower s<strong>pe</strong>ed than the fluid.<br />

Tournier and Judd (1945) reported that the s<strong>pe</strong>cific gravity of the ore is an important<br />

factor to consider. Heavier ores require more slo<strong>pe</strong> to be transported in an o<strong>pe</strong>n channel, as<br />

shown in Figure 6-14. Tournier and Judd (1945) reported that the size of the particles play<br />

an important role, and that larger particles require more slo<strong>pe</strong>, as shown in Figure 6-15.<br />

Figures 6-12 and 6-13 clearly demonstrate that it may be erroneous to use conventional<br />

Manning formulae for water flow de<strong>pe</strong>nding on the roughness of the pi<strong>pe</strong>, as these ignore<br />

the resultant roughness due to sand dunes and anti-dunes. Figures 6-14 and 6-15 clearly in-<br />

y m<br />

Flow direction<br />

wavelength<br />

liquid level<br />

FIGURE 6-13 Sand anti-dunes at high Froude number (Fr > 1).<br />

6.37


6.38 CHAPTER SIX<br />

s<strong>pe</strong>cific gravity = 3.8<br />

s<strong>pe</strong>cific gravity = 2.7<br />

slo<strong>pe</strong> of launder (%)<br />

0 20 40 60 80<br />

Weight concentration (%)<br />

FIGURE 6-14 The slo<strong>pe</strong> is a function of the s<strong>pe</strong>cific gravity of the ore as well as the weight<br />

concentration (after Tournier and Judd, 1945). The magnitude of the slo<strong>pe</strong> is not shown here as<br />

it de<strong>pe</strong>nds also on the hydraulic radius or sha<strong>pe</strong> of the launder.<br />

0 4 8 12 16 18 20<br />

particle size (mm)<br />

slo<strong>pe</strong> of launder (%)<br />

FIGURE 6-15 Larger particles require more slo<strong>pe</strong> (after Tournier and Judd, 1945). The<br />

magnitude of the slo<strong>pe</strong> is not shown here as it de<strong>pe</strong>nds also on the hydraulic radius or sha<strong>pe</strong> of<br />

the launder.


dicate that the density of the ore as well as the size of the solids do increase the slo<strong>pe</strong>. These<br />

important factors are unfortunately too often ignored by some engineers who rely on the<br />

Manning equation.<br />

6-6 FRICTION LOSSES AND SLOPE FOR<br />

HOMOGENEOUS SLURRY FLOWS<br />

Green et al. (1978) attempted to estimate the effect of particle sizes and concentration in<br />

the form of increased effective viscosity. In this approach, however, they essentially limited<br />

their studies to Newtonian slurries and did not take into account the effects of yield<br />

stress and plasticity, effects often encountered at high concentrations of fine particles.<br />

Geophysicists prefer to talk about cohesive bed forms when exploring the movement<br />

of clays in rivers. For certain soils, cohesive forces develop between the solid particles.<br />

Graf (1971) indicated that Equation 6-43 should be modified to include X0, a coefficient<br />

of cohesion of the material.<br />

For a Bingham <strong>slurry</strong>, the yield stress is added to Equation 6-45 to express the critical<br />

shear stress at the point of the incipient motion:<br />

�cr �� = � + X0 (6-70)<br />

(�s – �)gd� Cohesive (soil) materials include clay-sized (colloidal) particles, silt-sized particles,<br />

and sometimes sand-sized particles. Graf (1971) classified clays into the following three<br />

main categories:<br />

1. Kaolinites<br />

2. Montmorillonites<br />

3. Illites<br />

SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

6.39<br />

Other more minor clays include halloysites, chlorites, and vermiculites.<br />

Clay materials have residual electrostatic forces that attract cations and anions. This is<br />

measured as a cation exchange capacity in milliequivalents <strong>pe</strong>r 100 grams. Grim (1962)<br />

stated that Kaolinites have an exchange capacity of 3–15 milliequivalent <strong>pe</strong>r 100 grams,<br />

whereas illites rated higher at 10–40, and montmorillonites at 80–150.<br />

In very simple terms, Grim explained that there are two main structures for clays. One<br />

structure consists of two close sheets of packed hydroxyl molecules, in which aluminum,<br />

iron, and magnesium atoms are embedded in an octahedral coordination equidistant from<br />

the six oxygen atoms in the hydroxyls.<br />

The second structure consists of silica tetrahedrons. In the tetrahedron, the silicon<br />

atom is equidistant from the hydroxyls. The silica tetrahedral groups are arranged to form<br />

a hexagonal network, which is re<strong>pe</strong>ated from sheet to sheet.<br />

It would be beyond the sco<strong>pe</strong> of this book to discuss the physical as<strong>pe</strong>cts of clays. The<br />

<strong>slurry</strong> engineer should appreciate that these electrostatic forces and arrangement of chemical<br />

groups influence the yield stress at certain concentrations that may be at either the<br />

lower end or the higher end.<br />

Cohesive soils have the ability to absorb water and develop plasticity; they also have<br />

limit liquid. All three characteristics were briefly defined in Chapter 1. Cohesive colloids<br />

can form lumps. Forces on such lumps are shown in Figure 6-16.<br />

Graf (1971) summarized the test work of Karasev (1964) who proposed that the erosion<br />

of cohesive material beds in rivers is due to aggregate-to-aggregate rather than by


6.40 CHAPTER SIX<br />

Drag + Cohesion forces<br />

particle-to-particle contact. Karasev (1964) derived the following equation for the average<br />

scouring velocity:<br />

Ucr = 0.142 Ch� �1.2 + 8 �� 0.5<br />

2d50(�s – �L) + 3� ym ���<br />

(6-71)<br />

where<br />

� = information about cohesion<br />

Patm = atmospheric pressure<br />

A comparison between the computed and measured values of a scouring velocity is<br />

presented in Table 6-5. This value of critical s<strong>pe</strong>ed should be considered the s<strong>pe</strong>ed for<br />

minimum transportation of clay by hydrotransport.<br />

The flow of water in canals containing sand, cohesive soils, and cohesive banks was<br />

examined by Graf (1971) on the basis of the work of Simons et al. (1963). The graph in<br />

Figure 6-17 suggests that the hydraulic radius can be correlated to the flow rate by the following<br />

equation:<br />

RH = 0.43 · Q0.361 6-6-1 Bingham Plastics<br />

�L<br />

Lift Force<br />

Weight Force<br />

Whipple (1997) develo<strong>pe</strong>d numerical models for o<strong>pe</strong>n channel flow of Bingham fluids<br />

but did not provide a methodology to calculate friction losses. This pa<strong>pe</strong>r is important for<br />

a geophysicist but provides no useful tools for the <strong>slurry</strong> engineer.<br />

For a homogeneous <strong>slurry</strong>, there are two important numbers to calculate: the Reynolds<br />

number and the plasticity number (defined in Chapter 5). In the absence of well-defined<br />

models for friction losses of Bingham slurries, Abulnaga (1997) proposed a methodology<br />

to modify some of the equations of full-pi<strong>pe</strong> flows by expressing the Reynolds and Hedstrom<br />

numbers in terms of the hydraulic radius. At high shear rate, the coefficient of<br />

rigidity is taken as the viscosity for a Bingham plastic:<br />

Re = 4RHV�/� (6-72)<br />

� Patm<br />

Cohesion Force<br />

FIGURE 6-16 Forces on a colloidal lump moving in an o<strong>pe</strong>n channel.


SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

TABLE 6-5 Scouring Velocity of Clays (after Karasev, 1964)<br />

6.41<br />

Scouring velocity in streams<br />

______________________________<br />

Diameter of Karasev’s<br />

aggregate Adhesion � Ex<strong>pe</strong>rimental formula<br />

______________ _____________ _______________ _____________<br />

# Material mm in kPa psi m/s ft/sec m/s ft/sec<br />

1<br />

Aggregate materials<br />

Medium loam 2.2 0.087 225.6 32.7 1.74 3.28 1.83 6<br />

2 Rammed clay 3.0 0.118 550 80 2.16 7.09 2.82 9.3<br />

3 Rammed clay 4.0 0.157 202 29.3 2.06 6.76 2.36 7.74<br />

4 Heavy loam 2.7 0.106 373 54 2.04 6.69 2.22 7.28<br />

5 Heavy loam 4.0 0.157 225.6 32.7 1.20 3.94 1.70 5.58<br />

6 Clay 4.0 0.157 245 35.5 2.20 7.22 1.75 5.74<br />

7 Clay 4.0 0.157 285 41.3 1.30 4.26 1.86 6.1<br />

8 Clay 6.0 0.24 196 28.4 1.43 4.69 1.45 4.76<br />

9 Clay 4.0 0.157 285 41.3 1.54 5.05 1.86 6.10<br />

10 Compacted clay 0.8 0.03 510 73.9 2.23 7.32 3.20 1.05<br />

11 Heavy loam<br />

Dis<strong>pe</strong>rsed materials<br />

1.0 0.04 304 44.1 2.14 7.02 2.35 7.71<br />

12 Medium loam 4.0 0.157 746 108 1.91 6.27 2.88 9.45<br />

13 Clay 1.5 0.059 706 102 3.06 10.04 2.68 8.79<br />

14 Clay 3.2 0.126 785 114 2.35 7.71 2.4 7.87<br />

15 Clay 4.0 0.157 432 62.6 2.87 9.42 1.93 6.33<br />

16 Clay 0.8 0.03 785 114 2.40 7.87 2.42 7.94<br />

17 Clay 4.0 0.157 549 79.6 1.54 5.05 2.50 8.2<br />

FIGURE 6-17 Correlation between the hydraulic radius and the discharge flow rate. (From<br />

Graf, 1971, reprinted by <strong>pe</strong>rmission of McGraw-Hill.)


6.42 CHAPTER SIX<br />

And the Plasticity number is written in terms of the hydraulic radius as<br />

� 04R H<br />

PL = � (6-73)<br />

�V<br />

The Hedstrom number is the product of the Plasticity number and the Reynolds number<br />

and is calculated as<br />

16R H 2 ��0<br />

He = � (6-74)<br />

2 �<br />

In Chapter 3, the different categories of non-Newtonian flows were reviewed. Methods<br />

to calculate friction losses for Bingham slurries and power law slurries were presented<br />

in Chapter 5. Modified Reynolds and Hedstrom numbers for pseudoplastics were also<br />

introduced in Chapter 5.<br />

In Chapter 5, the Buckingham equation was introduced as<br />

He<br />

= – + (5-5)<br />

Modifying it for an o<strong>pe</strong>n channel in laminar flow yields<br />

4<br />

1 fNL He<br />

� � � � 2 3 8<br />

ReB 16 6 ReB 3 f NLReB � –<br />

� – (6-75)<br />

The Darby equation for the friction factor in turbulent regime is<br />

fNT = 10a 2 16RH��0/� � fNL �0 � � �2 4RHV� 16 6V �m<br />

b ReB 2<br />

��<br />

6(4RHV�m/�) 2<br />

� fNL � �<br />

4RHV�m 16<br />

where<br />

a = –1.47[1 + 0.146 exp(–2.9 × 10 –5 He)]<br />

b = –0.193<br />

The values of the parameters “a” and “b” were based on empirical data for closed conduits.<br />

They may be tentatively modified to o<strong>pe</strong>n channels to yield<br />

4R HV� m<br />

fNT = 10a� � b<br />

(5-10)<br />

where<br />

a = –1.47 [1 + 0.146 exp{–2.9 × 10 –5 2 (16RH �m�0/�2 )}]<br />

b = –0.193<br />

Equations for the empirical parameters “a” and “b” should be confirmed by extensive<br />

testing in o<strong>pe</strong>n channels. It is unfortunate that very little research is conducted in this extremely<br />

important field of fluid dynamics.<br />

Darby et al. (1992) proposed to combine the laminar and turbulent fanning friction<br />

factors into the following equation:<br />

fN = (f m NL + f m NT) (1/m) �<br />

�<br />

(5-11)<br />

where<br />

40,000<br />

m = 1.7 + � (5-12)<br />

ReB


SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

For an o<strong>pe</strong>n channel this becomes<br />

10,000�<br />

m = 1.7 + (6-76)<br />

� RHV� m<br />

6.43<br />

Example 6-12<br />

The soft high clay tailing from a cop<strong>pe</strong>r concentrator develops a Bingham viscosity of<br />

400 mPa · s at a weight concentration of 45% as well as a yield stress of 5 Pa. The <strong>slurry</strong><br />

density is 1350 kg/m3 . The tailings flow rate is 1600 m3 /hr. If the maximum allowed<br />

slo<strong>pe</strong> is 1.5%, determine the suitable size for a half-full smooth HDPE pi<strong>pe</strong> (i.e., ignore<br />

roughness factor).<br />

Solution in SI Units<br />

Iteration 1<br />

Assume a s<strong>pe</strong>ed of 1.8 m/s. Required area for flow is<br />

A = (1600/3600)/1.8 = 0.247 m2 Pi<strong>pe</strong> ID = [(8/�)A] 0.5 = 0.789 m (31.09 in)<br />

� = 0.4 Pa · s<br />

RH = DI/4 = 0.789/4 = 0.19725 � 0.198 m<br />

Re = 4RH�mV/� = 4 × 0.198 × 1350 × 1.8/0.4 = 2673<br />

He = 16 × 0.1982 × 1350 × 5/0.42 = 26,463<br />

fL � �1 + � � �1 + �<br />

fL � 0.0158<br />

a = –1.47[1 + 0.146 exp(– 2.9 × 10 –5Re)] a = –1.7<br />

fT = 10aRe0.193 = 0.00435<br />

m = 1.7 + 40,000/Re = 16.67<br />

m m 1/m fN = ( f L + f T ) = 0.016<br />

Using Equation 6-5:<br />

S =<br />

S = (0.016 × 1.82 fNV )/(2 × 9.81 × 0.198) = 0.0132<br />

S = 1.32%<br />

This approach was used by Abulnaga (1997) at Fluor Daniel Wright Engineers to design<br />

a tailings launder (see Figure 6-1) to transport tailings rich in soft high clay for a Peruvian<br />

cop<strong>pe</strong>r mine. The tailings system functioned well.<br />

The Wilson–Thomas method for full flow in closed channels does not rely on empirical<br />

coefficients such as the Darby method, but is based on the assumption that a thick sublayer<br />

lubricates the wall surface of the pi<strong>pe</strong>. It has not been modified yet for o<strong>pe</strong>n channel<br />

flows. This is a topic well worth further research.<br />

2<br />

16 He 16 26,463<br />

� � � �<br />

Re 6Re 2673 6 × 2673<br />

�<br />

2gRH


6.44 CHAPTER SIX<br />

6-7 FLOCCULATION LAUNDERS<br />

Hydraulic flocculation is sometimes conducted in launders feeding a thickener. The G<br />

gradient is a measure of the average shear rate for a flocculation tank. For a tank flocculated<br />

using a mixer, Camp (1955) defined the G gradient as:<br />

G = (6-77)<br />

��<br />

where<br />

P = power applied by the mixer<br />

Vol = volume of the liquid in the tank<br />

� = average viscosity of the solution<br />

Camp (1955) as well as the American Society of Civil Engineers (1977) modified this<br />

equation for launders, conduits, and flocculation basins by proposing that the power term<br />

P be replaced by the power due to friction in the launder:<br />

P = �Qg�H (6-78)<br />

where<br />

� = density of the mixture<br />

g = acceleration due to gravity<br />

Q = flow rate<br />

�H = head loss due to friction<br />

As a measure to control �H, ASCE proposed to install a baffle system in the flume.<br />

They defined this process as hydraulic flocculation:<br />

G = �� = (6-79)<br />

��<br />

Assuming that no baffles are installed in the launder, it would be possible to express<br />

the volume V as the product of the wetted area by the length of the launder:<br />

Vol = AL<br />

Assuming that the slo<strong>pe</strong> S of the launder is equal to the energy gradient (as <strong>pe</strong>r Equation<br />

6-5), or friction loss <strong>pe</strong>r unit length (or S = �H/L), then<br />

G = �� = P<br />

�<br />

Vol�<br />

P �Qg�H<br />

� �<br />

Vol� V�<br />

�Qg�H �QgS<br />

� � (6-80)<br />

V�A �� A�<br />

Equation 6-80 is valid only if no baffles or other artificial means are added to the launder<br />

to increase friction losses.<br />

Example 6-13<br />

A tailing is flowing to a thickener. In-line flocculation is applied to raise the viscosity to<br />

20 mPa · s. The flow rate is 1800 m3 /hr. The density of the mixture is 1420 kg/m3 . The<br />

launder is rectangular with a slo<strong>pe</strong> of 1.2%. The chemical process requires that the G gradient<br />

be smaller than 100. Determine a suitable size for the launder.<br />

Solution<br />

From Equation 6-79:<br />

100 = [1420(1800/3600) 9.81 × 0.012/(A × 0.02)] 0.5


SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

100 = 64.66 (1/A) 0.5<br />

A0.5 = 0.646 m<br />

A = 0.4179 m2 If the depth of the liquid is one-third the width of the launder,<br />

A = w2 /3<br />

W = 1.12 m<br />

Depth = 0.373 m<br />

Velocity = Q/A = 0.5 m3 /s/0.4179 m2 = 1.196 m/s<br />

6-8 FROUDE NUMBER AND STABILITY OF<br />

SLURRY FLOWS<br />

A measure of the stability of a flow in an o<strong>pe</strong>n channel can be expressed in the form of the<br />

Froude number (Fr), a ratio of inertia to gravity forces:<br />

V<br />

Fr = � (6-81)<br />

�g�ym�<br />

In the case of <strong>slurry</strong> flows, it is important to avoid subcritical flows (Fr < 0.8), as they<br />

often cause settling problems. Critical flows (0.8 < Fr < 1.5) may be associated with a degree<br />

of instability and wavy motion, leading in some cases to working problems and overflows.<br />

Kennedy (1963) reported test work on sand and suggested that antidunes occur at a<br />

Froude number of 0.8 to 1.4 (critical regime). Green et al. (1978) recommended that <strong>slurry</strong><br />

launders be designed for Froude numbers in excess of 1.5, to avoid regimes of instability<br />

of flow.<br />

On the other hand, excessively high Froude numbers (Fr > 5) are associated with<br />

steep slo<strong>pe</strong>s. Steep slo<strong>pe</strong> instability was discussed by Nie<strong>pe</strong>lt and Locher (1989). Slug<br />

flow is reported at high Froude numbers, causing working instability in the form of roll<br />

waves.<br />

Although economics may often dictate maintaining gentle slo<strong>pe</strong>s of < 2% on many<br />

long-distance tailings projects, the design of plants must too often accommodate tight<br />

spaces. The use of steep slo<strong>pe</strong>s (> 8%) may often cause high s<strong>pe</strong>eds, in excess of 6 m/s<br />

(20 ft/s). To avoid premature wear of liners or pi<strong>pe</strong>s, excessive slo<strong>pe</strong>s in plants should be<br />

avoided. If it is not possible to avoid steep slo<strong>pe</strong>s, drop boxes, pressurized boxes, chokes,<br />

or full-flow closed conduits should be used wherever wear is a major concern.<br />

There are certain conditions in an o<strong>pe</strong>n channel that may lead to a sudden change from<br />

a su<strong>pe</strong>rcritical regime to a subcritical regime. This is often associated with the so-called<br />

“hydraulic jump,” an increase in the depth of the liquid due to the lower s<strong>pe</strong>ed of motion.<br />

6-9 METHODOLOGY OF DESIGN<br />

6.45<br />

The design of o<strong>pe</strong>n launders is complex, with many implicit functions. Abulnaga (1997)<br />

suggested starting the calculations assuming a s<strong>pe</strong>ed of 2 m/s (6.56 ft/s). For many ty<strong>pe</strong>s<br />

of slurries, this is a good starting point. The flow rate is divided by the assumed s<strong>pe</strong>ed to


6.46 CHAPTER SIX<br />

obtain an area of the flow A. If a pi<strong>pe</strong> is selected, the flow area A is taken as � 45–55% of<br />

the cross-sectional area of commercial pi<strong>pe</strong>s to obtain a pi<strong>pe</strong> diameter.<br />

1. If a rectangular launder is to be manufactured, the width is used to compute the height<br />

of the liquid. For these launders, the height of the liquid is assumed to be 1/3 the width.<br />

2. For the area of the flow, the <strong>pe</strong>rimeter and hydraulic radius are then obtained.<br />

3. The Froude number is then calculated. If it ap<strong>pe</strong>ars that the flow will be in a subcritical<br />

or critical regime, the s<strong>pe</strong>ed should be increased.<br />

4. Re<strong>pe</strong>at steps 1 to 4 until the Froude number is larger than 1.5.<br />

5. Input the rheology of the <strong>slurry</strong> to obtain the plasticity, Reynolds, or Hedstrom number<br />

based on the hydraulic radius of the flow.<br />

6. Compute the friction factor in accordance with one of the equations listed.<br />

7. Obtain the friction loss <strong>pe</strong>r unit length and equate it to the slo<strong>pe</strong>, as <strong>pe</strong>r Equation 6-5.<br />

8. If the energy gradient or slo<strong>pe</strong> from Step 9 exceeds the physical contour of the terrain<br />

where the launder is to be installed, reiterate assuming a slower flow.<br />

9. Check on the deposition velocity or self-cleaning abilities of the solids. If the deposition<br />

velocity is more than 50% of the average velocity, s<strong>pe</strong>ed up the flow by changing<br />

the cross section of the launder or the physical slo<strong>pe</strong>.<br />

The following computer program, “Non-Newt-Channel,” uses the Darby method to<br />

design o<strong>pe</strong>n channel flow for non-Newtonian fluids on the basis of modifications to the<br />

Darby method.<br />

Computer Program “Non-Newt Channels”<br />

CLS<br />

PRINT “CHANNEL FLOW PROGRAM FOR NON-NEWTONIAN FLOWS<br />

PRINT “****************************************”<br />

PRINT<br />

c = 0<br />

pi = 4 * ATN(1)<br />

DEF fnlog10 (x) = LOG(x) * .43242944#<br />

DEF FNASN (x) = x + x ^ 3/6 + 3 * x ^ 5/(2 * 4 * 5) + 15 * x<br />

^ 7/(2 * 4 * 6 * 7) + (15 * 7)/(48 * 7 * 8) * x ^ 9<br />

DEF fnacos (x) = pi/2 – (x + x ^ 3/6 + (3/(2 * 4 * 5)) * x ^ 5 +<br />

15/(8 * 6 * 7) * x ^ 7 + 15 * 7 * x ^ 9/(48 * 7 * 8))<br />

g = 9.81<br />

INPUT “PROJECT “; proj$<br />

PRINT<br />

INPUT “DATE “; d$<br />

CLS<br />

INPUT “NAME OF ENGINEER “; e$<br />

‘e$<br />

CLS<br />

5 INPUT “AREA”; a$<br />

INPUT “LINE NUMBER “; li$<br />

CLS<br />

INPUT “DO YOU INTEND TO USE US UNITS (y/n)”; U$<br />

GOSUB CONVERSION<br />

INPUT “INITIAL FLOW RATE (m3/s)”; q<br />

‘INPUT “INITIAL FLOW RATE (m3/HR)”; QH<br />

‘q = QH/3600


SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

q = q * qul<br />

PRINT USING “flow rate = ##.### m3/s”; q<br />

PRINT<br />

INPUT “DESIGN FACTOR “; SF<br />

c = 0<br />

6 IF c = 1 THEN GOTO 19<br />

PRINT<br />

INPUT “s<strong>pe</strong>cific gravity of carrier liquid”; sgl<br />

INPUT “s<strong>pe</strong>cific gravity of solids”; sgs<br />

PRINT<br />

PRINT “ please choose between input of weight or volume<br />

concentration”<br />

PRINT “ 1- weight concentration”<br />

PRINT “ 2- volume concentration”<br />

PRINT<br />

12 INPUT “ 1 or 2”; cwe<br />

IF cwe = 1 THEN INPUT “weight concentration in <strong>pe</strong>rcent”; cwin<br />

IF cwe = 2 THEN INPUT “volume concentration in <strong>pe</strong>rcent”; cvin<br />

IF cwe = 0 OR cwe > 2 THEN GOTO 12<br />

PRINT<br />

IF cwe = 1 THEN cw = cwin/100<br />

IF cwe = 1 THEN sgm = sgl/(1 - cw * (1 - sgl/sgs))<br />

IF cwe = 1 THEN cv = (sgm - sgl)/(sgs - sgl)<br />

IF cwe = 1 THEN cvin = 100 * cv<br />

IF cwe = 1 THEN PRINT USING “s<strong>pe</strong>cific gravity of mixture = ##.##,<br />

cv = #.###”; sgm; cv<br />

IF cwe = 2 THEN cv = cvin/100<br />

IF cwe = 2 THEN sgm = cv * (sgs - sgl) + sgl<br />

IF cwe = 2 THEN cw = cv * sgs/sgm<br />

IF cwe = 2 THEN cwin = cw * 100<br />

IF cwe = 2 THEN PRINT USING “s<strong>pe</strong>cific gravity of mixture = ##.##,<br />

cwin = ##.##%”; sgm; cw<br />

dens = sgm * 1000<br />

PRINT<br />

INPUT “DO YOU KNOW THE VISCOSITY (y/n)”; ZS$<br />

IF ZS$ = “Y” OR ZS$ = “y” THEN INPUT “VISCOSITY (mPa.s)”; vu1<br />

CLS<br />

IF ZS$ = “N” OR ZS$ = “n” THEN KRAT = 1 + 2.5 * cv + 10.05 * cv ^ 2<br />

+ .00273 * EXP(16.6 * cv)<br />

‘ASSUMED VISCOSITY OF WATER 1 mPa · s<br />

IF ZS$ = “N” OR ZS$ = “n” THEN vu = KRAT * .001<br />

IF ZS$ = “Y” OR ZS$ = “y” THEN vu = vu1/1000<br />

CLS<br />

PRINT USING “VISCOSITY = ##.##### Pa.s”; vu<br />

INPUT “hit any key”; t$<br />

CLS<br />

INPUT “yield stress in dynes/cm2 “; y1<br />

y = y1/10<br />

CLS<br />

PRINT “you can let the program iterate for itself or you can input<br />

an initial s<strong>pe</strong>ed”<br />

PRINT<br />

v1 = 2<br />

PRINT “iteration starts at 2 m/s (6.6 ft/s)”<br />

INPUT “do you prefer to input an initial s<strong>pe</strong>ed (Y/N)”; b$<br />

6.47


6.48 CHAPTER SIX<br />

IF b$ = “N” OR b$ = “n” THEN GOTO 18<br />

17 PRINT USING “ current s<strong>pe</strong>ed = ##.## m/s “; v1<br />

IF us = 1 THEN INPUT “ initial s<strong>pe</strong>ed in ft/s “; v1<br />

IF us = 2 THEN INPUT “ initial s<strong>pe</strong>ed in m/s “; v1<br />

v1 = v1 * leg<br />

18 INPUT “maximum allowed slo<strong>pe</strong> in <strong>pe</strong>rcent “; s1<br />

19 IF c = 1 THEN q = q * SF<br />

IF c = 1 THEN GOTO 50<br />

21 PRINT “choose sha<strong>pe</strong> of launder from following list “<br />

PRINT<br />

PRINT “1- rectangular “<br />

PRINT “2- circular”<br />

PRINT “3- U sha<strong>pe</strong>”<br />

INPUT “which choice (1,2,3 etc.....)”; ck<br />

50 IF ck = 1 THEN GOSUB rect<br />

IF ck = 2 THEN GOSUB circ<br />

IF ck = 3 THEN GOSUB usha<strong>pe</strong><br />

PRINT “v1”<br />

re = v1 * 4 * rh * dens/vu<br />

he = (4 * rh/vu) ^ 2 * dens * y<br />

PRINT USING “Reynolds No = #########, Hedstrom No = ######## “; re;<br />

he<br />

GOSUB friction<br />

IF c = 0 THEN GOSUB settling<br />

IF v2m > (v1/2) THEN PRINT “to avoid settling, flow should be<br />

s<strong>pe</strong>eded up”<br />

IF v2m > (v1/2) THEN GOSUB increase<br />

IF sm > s1 THEN PRINT “slo<strong>pe</strong> exceeds maximum allowed slo<strong>pe</strong>”<br />

IF sm > s1 THEN GOTO 17<br />

‘INPUT “do you want to print out these results as a minimum slo<strong>pe</strong>”;<br />

min$<br />

‘IF min$ = “Y” OR min$ = “y” THEN min = 1<br />

‘IF min = 1 THEN GOSUB print1<br />

‘IF min = 1 THEN GOTO 999999<br />

GOSUB gradient<br />

INPUT “do you want a hard copy (Y/N)”; mv$<br />

IF mv$ = “y” OR mv$ = “Y” THEN GOSUB print1<br />

999999 IF SF > 1 THEN c = c + 1<br />

IF c = 1 THEN GOTO 19<br />

END<br />

increase:<br />

v1 = v1 * 1.03<br />

PRINT USING “velocity = ##.## m/s”; v1<br />

area = q/v1<br />

RETURN<br />

decrease:<br />

v1 = v1 * .97<br />

area = q/v1<br />

RETURN<br />

are:<br />

area = q/v1<br />

PRINT “flow area “; area<br />

RETURN


SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

CONVERSION:<br />

IF U$ = “Y” OR U$ = “y” THEN us = 1<br />

IF U$ = “n” OR U$ = “N” THEN us = 2<br />

IF us = 1 THEN PRINT “for us units use foot for length, US gallon<br />

for flow”<br />

IF us = 1 THEN PRINT “s<strong>pe</strong>ed in ft/s”<br />

ft = .3048<br />

gal = 3.785<br />

inch = .0254<br />

IF us = 2 THEN GOTO 888<br />

leg = ft<br />

qul = gal/60000<br />

GOTO 890<br />

888 leg = 1<br />

qul = 1<br />

890<br />

RETURN<br />

6.49<br />

rect:<br />

CLS<br />

IF c = 1 THEN GOTO 1600<br />

PRINT “you have chosen a rectangular launder - do you want to<br />

continue (Y/N)”; mre$<br />

IF mre$ = “n” OR mre$ = “N” THEN GOTO 21<br />

14 IF us = 1 THEN INPUT “width of channel (ft) “; w<br />

IF us = 2 THEN INPUT “width of channel (m) “; w<br />

IF w = 0 THEN GOTO 14<br />

w = w * leg<br />

PRINT<br />

INPUT “do you want the program to calculate height of walls (Y/N)”;<br />

hr$<br />

IF hr$ = “N” OR hr$ = “n” THEN INPUT “height of walls “; hl<br />

IF hr$ = “y” OR hr$ = “Y” THEN hl = .5 * w<br />

hl = hl * leg<br />

1600 area = q/v1<br />

dep = area/w<br />

IF c = 1 THEN GOTO 1606<br />

IF p$ = “y” OR p$ = “Y” THEN GOTO 1606<br />

INPUT “what ratio of fill is acceptable (0.333, 0.5, 0.75)”; fill<br />

1605 IF hr$ = “N” OR hr$ = “n” THEN hl = dep/fill<br />

hfill = hl * fill<br />

1606 IF (dep > hfill) THEN PRINT “depth of liquid exceeds preferred<br />

fill ratio”<br />

IF dep > hfill THEN v1 = v1 * 1.01<br />

IF dep > hfill THEN area = q/v1<br />

IF dep > hfill THEN dep = area/w<br />

IF dep > hfill THEN GOTO 1605<br />

‘calculation of hydraulic radius<br />

rh = area/(w + 2 * dep)<br />

mhd = dep<br />

GOSUB froude<br />

IF nf < = 1.5 THEN PRINT “froude number too low at “; nf<br />

IF nf < = 1.5 THEN INPUT “flow is unstable do you want to stabilize<br />

the flow “; p$<br />

IF p$ = “N” OR p$ = “n” THEN GOTO 1612<br />

IF nf > 1.5 THEN GOTO 1612<br />

v1 = v1 * 1.01


6.50 CHAPTER SIX<br />

area = q/v1<br />

GOTO 1600<br />

1612 RETURN<br />

circ:<br />

rf = 0<br />

PRINT “calculation for a circular launder”<br />

‘RINT “ITERATION STARTS FOR 1/2 FULL PIPE”<br />

INPUT “degree of fullness as a ratio of area of flow to pi<strong>pe</strong> area<br />

(.5,.6 etc..)”; full1<br />

1999 PRINT “s<strong>pe</strong>ed (m/s)”; v1<br />

INPUT “do you want to change s<strong>pe</strong>ed (Y/N)”; lk$<br />

IF lk$ = “y” OR lk$ = “Y” THEN INPUT “new s<strong>pe</strong>ed in m/s “; v1<br />

‘ IF lk$ = “n” OR lk$ = “N” THEN GOTO 2001<br />

2095 area = q/v1<br />

IF c = 1 THEN GOTO 456<br />

dia = SQR(area * 4/(full1 * pi))<br />

IF us = 1 THEN diapus = dia/.0254<br />

IF us = 1 THEN PRINT USING “recommended inner pi<strong>pe</strong> diameter = ##.##<br />

in “; diapus<br />

IF us = 2 THEN PRINT USING “recommended inner pi<strong>pe</strong> diameter =<br />

###.####m”; dia<br />

2001<br />

IF us = 2 THEN GOTO 2100<br />

IF nff = 0 THEN GOTO 2002<br />

PRINT USING “present pi<strong>pe</strong> i.d = ##.### m”; id<br />

IF us = 1 THEN idus = id/.0254<br />

IF us = 1 THEN PRINT “present pi<strong>pe</strong> id = ###.##inches”; idus<br />

2002 INPUT “ pi<strong>pe</strong> outer diameter in inches”; dout<br />

IF rf > 0 THEN GOTO 2004<br />

INPUT “pi<strong>pe</strong> thickness in inches “; thickus<br />

INPUT “pi<strong>pe</strong> liner thickness in inches “; linus<br />

2004 idin = dout - 2 * thickus - 2 * linus<br />

PRINT USING “pi<strong>pe</strong> i.d = ###.### in “; idin<br />

id = idin/12<br />

PRINT USING “pi<strong>pe</strong> id = ##.## ft”; id<br />

GOTO 2105<br />

2100 INPUT “pi<strong>pe</strong> outer diameter in mm”; d2m<br />

IF nff = 1 THEN GOTO 2101<br />

IF rf > 1 THEN GOTO 2101<br />

INPUT “pi<strong>pe</strong> thickness in mm”; thick<br />

INPUT “pi<strong>pe</strong> liner thickness in mm”; lin<br />

2101 idm = d2m - 2 * thick - 2 * lin<br />

id = idm/1000<br />

2105 id = id * leg<br />

r1 = id/2<br />

a2 = pi * r1 ^ 2<br />

456 IF area < a2 THEN GOSUB depth1<br />

IF area > a2 THEN GOSUB depth2<br />

IF a2 = area THEN PRINT “DEPTH = RADIUS”<br />

RATIO1 = dep/(2 * r1)<br />

INPUT “HIT ANY KEY TO CONTINUE”; l$<br />

RATIO1 = dep/id<br />

457 PRINT “RATIO OF LIQUID DEPTH TO DIAMETER”; RATIO1<br />

IF RATIO1 < 1.05 * full1 AND RATIO1 > .95 * full1 THEN GOTO 470<br />

IF RATIO1 < .2 THEN GOTO 490


SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

IF RATIO1 > .85 THEN GOTO 492<br />

IF RATIO1 < .48 THEN INPUT “ DO YOU WANT TO INCREASE THE DEPTH OF<br />

LIQUID TO REDUCE SLOPE (y/n)”; n$<br />

IF RATIO1 > .52 THEN INPUT “do you want to decrease liquid depth “;<br />

dp$<br />

IF RATIO1 < .48 AND n$ = “N” OR n$ = “n” THEN GOTO 470<br />

IF RATIO1 < .48 AND n$ = “y” OR n$ = “y” THEN GOTO 465<br />

IF RATIO1 > .52 AND dp$ = “N” OR dp$ = “n” THEN GOTO 470<br />

IF RATIO1 > .52 AND dp$ = “Y” OR dp$ = “y” THEN GOTO 467<br />

GOTO 470<br />

490 PRINT “ please reduce pi<strong>pe</strong> diameter as depth is less than 20%<br />

of diameter”<br />

INPUT “do you want to decrease the pi<strong>pe</strong> diameter “; kjl$<br />

IF kjl$ = “n” OR kjl$ = “N” THEN GOTO 456<br />

IF kjl$ = “Y” OR kjl$ = “y” THEN GOTO 2001<br />

492 PRINT “please increase pi<strong>pe</strong> diameter as depth is more than 90%<br />

of diameter”<br />

INPUT “do you want to change pi<strong>pe</strong> diameter (Y/N)”; qg$<br />

IF qg$ = “Y” OR qg$ = “y” THEN GOTO 465<br />

IF qg$ = “n” OR qg$ = “N” THEN GOTO 470<br />

465 GOSUB decrease<br />

GOSUB are<br />

GOSUB depth1<br />

GOTO 456<br />

467 GOSUB increase<br />

GOSUB are<br />

GOSUB depth2<br />

GOTO 456<br />

470 PRINT “s<strong>pe</strong>ed (m/s)”; v1<br />

RATIO1 = dep/id<br />

INPUT “hit any key to continue”; l$<br />

GOSUB angle<br />

2120 mhd = dep<br />

mhd = id * (fnacos(1 - 2 * RATIO1) - (2 - 4 * RATIO1) * SQR(ABS<br />

(RATIO1 - RATIO1 ^ 2)))/(8 * SQR(ABS(RATIO1 - RATIO1 ^ 2)))<br />

mhds = mhd/.0254<br />

PRINT USING “mean hydraulic depth = ##.##m ##.### in”; mhd; mhds<br />

GOSUB froude<br />

PRINT “FROUDE NUMBER”, nf<br />

INPUT “HIT ANY KEY TO CONTINUE”; lk$<br />

IF nf < = .8 THEN PRINT “a new diameter is recommended”<br />

IF nf < = 1.4 THEN GOTO 1999<br />

IF nf < = .8 THEN GOSUB increase<br />

IF nf < = .8 THEN GOSUB are<br />

IF nf < = .8 THEN PRINT “flow is subcritical”<br />

3000 IF (nf > .8) AND (nf < 1.5) THEN GOSUB increase<br />

IF (nf > .8) AND (nf < 1.5) THEN GOSUB are<br />

IF (nf > .8) AND (nf < 1.5) THEN PRINT “flow is critical”<br />

IF nf < 1.5 THEN nff = 1<br />

IF nf > = 1.5 THEN nff = 0<br />

IF nf < 1.5 THEN GOTO 456<br />

30001 GOSUB angle<br />

PRINT “<strong>pe</strong>rimeter”; <strong>pe</strong>r<br />

PRINT “area “; area<br />

rh = area/<strong>pe</strong>r<br />

PRINT “hydraulic radius”; rh<br />

6.51


6.52 CHAPTER SIX<br />

INPUT “HIT ANY KEY TO CONTINUE”; l$<br />

RETURN<br />

usha<strong>pe</strong>:<br />

RETURN<br />

froude:<br />

nf = v1/SQR(g * mhd)<br />

IF nf < .8 THEN PRINT “flow is subcritical”<br />

IF (nf > .8) AND (nf < 1.5) THEN PRINT “flow is critical”<br />

PRINT “froude number = “; nf<br />

RETURN<br />

friction:<br />

a = -1.378 * (1 + .146 * EXP(–.000029 * he))<br />

PRINT “reynolds “; re<br />

m = 1.7 + 40000/re<br />

PRINT USING “factor a = ###.###### and exponent m = ##.###”; a; m<br />

PRINT<br />

INPUT “hit any key to continue “; kkkkkkk$<br />

FTU = (10 ^ a) * re ^ (-.193)<br />

PRINT “ft = “; FTU<br />

PRINT<br />

fl = (16/re) * (1 + he/(6 * re))<br />

PRINT “fl = “; fl<br />

ff = (fl ^ m + FTU ^ m) ^ (1/m)<br />

fd = 4 * ff<br />

IF c > 1 THEN GOTO 666<br />

PRINT USING “in absence of roughness fanning = #.###### and darcy =<br />

#.######”; ff; fd<br />

[A section of the program here lists all ty<strong>pe</strong>s of materials and<br />

their roughness as explained by table 6-2, it is not reproduced<br />

here to save space<br />

em refers to absolute roughness in meters and emf in ft]<br />

PRINT USING “estimated roughness for new system = ##.##### m ##.###<br />

ft”; em; emf<br />

666 FOR i = 1 TO 20<br />

fd2 = fd<br />

ro = (em/(3.7 * 4 * rh) + 2.51/(re * SQR(fd)))<br />

h = -2 * fnlog10(ro)<br />

fd = h ^ -2<br />

NEXT i<br />

dg = fd2 - fd<br />

PRINT “revised darcy factor to account for roughness”; fd<br />

PRINT<br />

PRINT “iteration error on darcy “; dg<br />

ch2 = SQR(8 * g/fd)<br />

n2 = rh ^ (1/6)/ch2<br />

PRINT USING “Chazy No = ###.## and Manning number = #.#####<br />

(including roughness)”; ch2; n2<br />

s2 = fd * v1 ^ 2/(8 * rh * 9.81)<br />

sm = s2 * 100<br />

PRINT USING “recommended slo<strong>pe</strong> = ##.### % “; sm<br />

PRINT<br />

RETURN<br />

settling:<br />

REM check for any coarse particles being transported in a Non-New-


SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

tonian mixture<br />

PRINT “iteration on settling s<strong>pe</strong>ed for particles using Camp<br />

equation”<br />

INPUT “particle size (mm) “; dp<br />

dp2 = .001 * dp/ft<br />

v2 = SQR((8 * .8 * 32 * dp2 * (dens/1000 - 1))/fd)<br />

v2m = v2 * ft<br />

PRINT USING “SETTLING SPEED = #.## m/s ##.## ft/s”; v2m; v2<br />

IF v1 < (v2m * 2) THEN PRINT “warning settling s<strong>pe</strong>ed is higher than<br />

half of average s<strong>pe</strong>ed”<br />

RETURN<br />

gradient:<br />

‘grad = (2 * vu/dens) ^ (–.5) * (((fd/(4 * rh)) ^ .5) * v1 ^ 1.5)<br />

grad = (dens * q * 9.81 * s2/(area * vu)) ^ .5<br />

PRINT USING “velocity gradient = ###.## sec-1”; grad<br />

RETURN<br />

depth1:<br />

d2 = .1 * r1<br />

777 LE = r1 - d2<br />

beta = fnacos(LE/r1)<br />

PRINT “angle beta”; beta<br />

‘INPUT “hit any key to continue”; lllll$<br />

A3 = r1 ^ 2 * (beta - SIN(beta) * COS(beta))<br />

IF A3 < (.975 * area) THEN d2 = d2 + .01 * r1<br />

IF A3 < (.975 * area) THEN GOTO 777<br />

IF A3 > (1.025 * area) THEN dpf = 1<br />

IF A3 > (1.025 * area) THEN GOSUB depth2<br />

PRINT “DEPTH OF SLURRY”; d2<br />

dep = d2<br />

‘INPUT “hit any key to continue”; k$<br />

RETURN<br />

depth2:<br />

IF dpf = 1 THEN GOTO 778<br />

d2 = .9 * r1<br />

778 LE = d2 - r1<br />

beta = FNASN(LE/r1)<br />

REM next line changed for rev 1.02 - pi in front of beta removed<br />

A3 = pi * r1 ^ 2/2 + beta * r1 ^ 2 + r1 ^ 2 * SIN(beta) * COS(beta)<br />

IF A3 > 1.025 * area THEN d2 = d2 - .01 * r1<br />

IF A3 > 1.025 * area THEN GOTO 778<br />

IF A3 < .975 * area THEN GOSUB depth1<br />

dep = d2<br />

depus1 = dep/.0254<br />

PRINT USING “depth = ##.### m ###.### in”; dep; depus1<br />

INPUT “hit any key to continue”; k$<br />

RETURN<br />

angle:<br />

IF dep < r1 THEN theta = fnacos((dep - r1)/r1)<br />

IF dep > r1 THEN theta = F NASN((dep - r1)/r1)<br />

IF dep = r1 THEN theta = pi/2<br />

IF dep < r1 THEN theta = 2 * theta<br />

IF dep > r1 THEN theta = 2 * theta + pi<br />

<strong>pe</strong>r = theta * r1<br />

RETURN<br />

6.53


6.54 CHAPTER SIX<br />

Flow may accelerate at bends due to the formation of centrifugal forces. The velocity<br />

profile is then distorted (Einstein and Hardner, 1954).<br />

6-10 SLURRY FLOW IN CASCADES<br />

Cascades are important mechanisms for the transportation of <strong>slurry</strong>. They are steep o<strong>pe</strong>n<br />

channels and are associated with a high Froude number and steep gradients. Stricklen<br />

(1984) suggested that cascades be used on slo<strong>pe</strong>s between 5% and 65% with velocities in<br />

excess of 10 m/s (33 ft/sec). At these magnitudes of s<strong>pe</strong>ed, excessive wear would occur<br />

on the walls of the o<strong>pe</strong>n channel cascade.<br />

There are three ty<strong>pe</strong>s of boxes to consider for reducing the s<strong>pe</strong>ed:<br />

1. Cascade feed box (Figure 6-18)<br />

2. Cascade receiving sump (Figure 6-19)<br />

3. Siphon feed box (Figure 6-20)<br />

Stricklen (1984) suggested that under certain conditions the localized solid concentration<br />

may exceed 65% by volume and may cause a pattern of “slug” flow with considerable<br />

localized wear. To mitigate against this problem, while controlling the s<strong>pe</strong>ed, he suggested<br />

that the launder be designed as wide as possible to reduce the hydraulic radius and<br />

depth of the flow, but still narrow enough as to avoid slug flow.<br />

Two parameters need to be computed in order to check for localized slug flow.<br />

1. The Vedernikov number Ve:<br />

U<br />

Ve = ��<br />

(6-82)<br />

(gym cos �) 1/2<br />

2 bw � �<br />

3 Pw<br />

Worn-out mill liner<br />

used to absorb wear<br />

Worn-out pump liner<br />

used to absorb wear<br />

Fig 6-19<br />

Low entry slo<strong>pe</strong><br />

Side ventilation window<br />

(recommended for deep drops)<br />

Nap<strong>pe</strong> of <strong>slurry</strong><br />

D<br />

Minimum D/3<br />

Steep outlet cascade<br />

FIGURE 6-18 Entry into a cascade feed box from a low-slo<strong>pe</strong> launder.


worn out mill liner<br />

used to absorb wear<br />

worn out pump liner<br />

used to absorb wear<br />

feed pi<strong>pe</strong><br />

SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

steep cascade at inlet<br />

FIGURE 6-20 Siphon feed pi<strong>pe</strong> drop box.<br />

nap<strong>pe</strong> of <strong>slurry</strong><br />

(ventilation window not shown)<br />

low slo<strong>pe</strong> for outlet launder<br />

FIGURE 6-19 Entry into a cascade receiving sump from a steep launder.<br />

Fig 6-21<br />

pi<strong>pe</strong> tee fitting<br />

6.55<br />

discharge pi<strong>pe</strong><br />

2. The Montuori number M:<br />

M 2 = (6-83)<br />

where<br />

bw = bottom width of the channel<br />

Pw = wetted <strong>pe</strong>rimeter of the channel<br />

� = tan –1 (h/L) = tan –1 U<br />

S<br />

L = length of the channel<br />

Figure 6-21 shows a linear limit between the Vedernikov and the Montuori numbers.<br />

Below the line, no slug flow occurs and the flow is stable. Above the line, slug flow occurs.<br />

2<br />

��<br />

gSL cos �


6.56 CHAPTER SIX<br />

FIGURE 6-21 The correlation between the Vedernikov number and the square of the Montuori<br />

number squared is used to differentiate between slug and no-slug flows. (From Stricklen,<br />

1984.)<br />

If the calculations of the Vedernikov and Montuori numbers indicate that the flow is<br />

of a slug ty<strong>pe</strong>, it will be necessary to determine the intermediate points from which unstable<br />

rolling waves would be generated.<br />

Nie<strong>pe</strong>lt and Locher (1989) as well as Stricklen (1984) proposed to compute a sha<strong>pe</strong><br />

factor for the chute:<br />

x =<br />

y m<br />

� Pw<br />

where<br />

Pw = wetted <strong>pe</strong>rimeter<br />

ym = average depth of the <strong>slurry</strong> in the channel<br />

Steep launders may cause the formation of roll-waves that are associated with instability.<br />

The Vedernikov number may be used as a design guide to determine these areas. Nie<strong>pe</strong>lt<br />

and Locher (1989) extended the analysis to slurries and showed a marked difference with<br />

water flows (Figure 6-22).<br />

6-11 HYDRAULICS OF THE DROP BOX AND<br />

THE PLUNGE POOL<br />

Certain remote mines in mountainous regions have chosen over the years to dispose of<br />

their tailings at sea level and sometimes to submerge them in the sea. The drop box has<br />

been found to be an effective method to achieve energy dissipation during transportation.<br />

There are particular design criteria that the drop box or receiving sump must meet to<br />

avoid rapid wear of its walls:


6.57<br />

1<br />

� M 2<br />

= gsL cos (�)<br />

��<br />

U 2<br />

FIGURE 6-22 The Vedernikov number is used as a design guide to determine roll waves associated with steep cascades. There is,<br />

however, a marked difference between water and slurries. (From Nie<strong>pe</strong>lt and Locher, 1989, reprinted by <strong>pe</strong>rmission of SME.)


6.58 CHAPTER SIX<br />

� The incoming liquid or nap<strong>pe</strong> should impact the <strong>slurry</strong> liquid surface in the drop box<br />

and not the bottom surface or walls.<br />

� The sump should be sized sufficiently large for its walls to be outside the computed<br />

area of impingement or high turbulence.<br />

� If slug flows or flows at high Froude numbers are allowed to enter the receiving sump,<br />

the sump should be fairly long to co<strong>pe</strong> with the fluctuations of flows.<br />

� A weir may be installed in the receiving sump to reduce the length of the hydraulic<br />

jump.<br />

� Froth arresters are recommended for frothy slurries.<br />

� The area of high turbulence or the exit from the receiving sump may have to be covered<br />

to avoid overfills.<br />

The design of such sumps is far from easy. In the next section, the mathematics of the<br />

<strong>slurry</strong> fall will be presented to the reader in a brief practical approach. Excellent books on<br />

the engineering of small dams are available for further reading.<br />

One question often asked is what is the recommended depth of a plunge pool. The<br />

rule of thumb in the case of water is that the plunge pool should be one-third the depth<br />

of the waterfall. That means that for a waterfall drop of 30 m one would need to provide<br />

an additional depth of 10 m to absorb all the turbulence. This is not always possible<br />

to achieve, and energy dissipaters are then introduced to absorb the turbulence. In<br />

mining, these energy dissipaters are often worn-out mill liners, pump liners, or im<strong>pe</strong>llers<br />

that are put at the bottom of the plunge pool to wear away as they absorb the impact of<br />

abrasive <strong>slurry</strong> fall.<br />

In this chapter, we shall consider the more common drop box found in many mining<br />

plants. The economics and the size of many projects, as well as wear considerations, often<br />

reduce the problem to rectangular or circular drop boxes. Other forms of energy dissipaters<br />

such as ogees and ski jumps that are discussed in certain books on civil engineering have<br />

not found application in mining because of the problem of lining such complex sha<strong>pe</strong>s.<br />

For a rectangular entry into the fall, the analysis of this problem is based on dividing<br />

flow rate Q by the width of the launder before the fall:<br />

Q<br />

qb = � (6-84)<br />

w<br />

The following analysis assumes a constant width of the launder starting well upstream<br />

from the fall.<br />

If y is the depth of the liquid well upstream of the fall, and V is the velocity of the liquid,<br />

as in Figure 6-23, the total energy is<br />

V<br />

H = y + (6-85)<br />

If the flow is subcritical well upstream from the fall, it will tend to accelerate near the<br />

fall. Rubin (1997) demonstrated that the minimum energy head for a waterfall occurs<br />

when the flow prior to the drop is in a critical regime with a Froude number of 1.0. Under<br />

such conditions, the flow accelerates toward the brink of the fall, thus reducing the depth<br />

Yb, which according to Fathy and Shaarawi (1954) would be<br />

2<br />

�<br />

2g<br />

Y b<br />

� Y0<br />

= 0.716 (6-86)


Y<br />

subcritical flow<br />

SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

flow <strong>pe</strong>r unit length q =Q/b<br />

b<br />

Total Energy Line<br />

2<br />

(V /2g)<br />

3<br />

2<br />

Y = Q /b<br />

0<br />

5 Y 0<br />

The Froude number of 1.0 occurs five times the critical depth upstream of the brink. The<br />

critical depth is defined as<br />

Y0 = � � 1/3<br />

(6-87)<br />

For water flow, the critical slo<strong>pe</strong> is expressed in terms of the critical depth and the<br />

Manning roughness factor as<br />

1<br />

S0 = � (6-88)<br />

[Y0] 1/3<br />

gn2 �<br />

g<br />

�<br />

Fr<br />

But since Fr = 1.0, Equation 6-88 is also expressed as<br />

S 0 =<br />

q b 2<br />

gn 2<br />

� [Y0] 1/3<br />

6.59<br />

Obviously, for slurries with different roughness values due to the deposition of sediments<br />

or formation of antidunes, Equation (6-88) is not readily applicable.<br />

From the point of view of the designer of a <strong>slurry</strong> drop box, it is important to determine<br />

the area of impingement of the jet, the depth of the backwater, and the area of the<br />

still water, in order to provide pro<strong>pe</strong>r liners and protection from wear. The nap<strong>pe</strong> must be<br />

pro<strong>pe</strong>rly ventilated, as in Figure 6-24; otherwise the <strong>slurry</strong> may tear the structure apart.<br />

It may ap<strong>pe</strong>ar strange to the reader that the author is focusing on the case of minimum<br />

energy with entry in a subcritical flow, although we have reiterated in previous sections of<br />

this chapter the need to maintain a su<strong>pe</strong>rcritical flow for slurries in launders. The minimum<br />

energy entry is a case of reference used to understand more complex flows at high<br />

Froude number in which the projection of the nap<strong>pe</strong> is even further away. There are cases<br />

in which entry is at minimum energy, such as from a lake into a river, or from a large tailings<br />

pond into an o<strong>pe</strong>n channel, or from a relatively horizontal channel into a large drop<br />

box used for sampling the tailings. In fact, entering the fall at minimum energy allows for<br />

a better capture of samples for analysis (Figure 6-25).<br />

Y 0<br />

flow Q<br />

VENTILATION AIR<br />

FIGURE 6-23 Entering a waterfall with minimum energy gradient.<br />

width "b"<br />

Y = 0.716


6.60 CHAPTER SIX<br />

FIGURE 6-24 This drop box for a large tailings flow features three 24� ventilation windows<br />

in each side wall to <strong>pe</strong>rmit ventilation under the nap<strong>pe</strong>.<br />

The energy dissipation at the bottom of the fall was discussed in detail by Moore<br />

(1943) and Rand (1955). The hydraulics of such a fall will therefore be summarized here<br />

for practical design considerations, with focus on the main equations.<br />

Rand observed three different flows for a waterfall with a well-ventilated nap<strong>pe</strong>,<br />

which are depicted in Figures 6-26 to 6-27. In the first case, Case A (Figure 6-27), the<br />

flow approaches the crest of the waterfall in a subcritical regime. The flow is characterized<br />

by a nonsubmerged nap<strong>pe</strong> at the point of impingement with the apron. Rand indicated<br />

without definite proof that the height of the liquid at the crest is 0.715 of the critical<br />

depth. The region between the wall and the nap<strong>pe</strong> is called the under-nap<strong>pe</strong>. It has a depth<br />

d f which is higher than the flow downstream of the point of impingement. In the undernap<strong>pe</strong>,<br />

the flow is recirculating.<br />

As the nap<strong>pe</strong> hits the apron, it turns smoothly into su<strong>pe</strong>rcritical regime at a distance L d<br />

from the wall. This distance L d is called the drop distance. At the point of impingement,<br />

the depth of the stream reaches a minimum with a depth d 1 at L d from the wall. After d 1,<br />

the flow depth increases smoothly while remaining in a su<strong>pe</strong>rcritical regime until a certain<br />

distance L j and a depth d b, where a stationary hydraulic jump occurs between the su<strong>pe</strong>rcritical<br />

and subcritical flows. The depth of the flow increases until a steady level is<br />

reached, d 3, called the tail water depth.<br />

Case B (Figure 6-28) is described by Rand as a borderline case. By comparison with<br />

Case A, the flow is critical or slightly su<strong>pe</strong>rcritical before the crest of the fall. There is no<br />

relative distance between d 1 and d 3, and the hydraulic jump occurs practically at the region<br />

of the impingement with the apron and extends over a distance L until a steady-state<br />

d 2 is reached for the tail water. The nap<strong>pe</strong> is not submerged, but there is no su<strong>pe</strong>rcritical<br />

flow over the apron, so the distance between the region of impingement and the tail water<br />

is considered the shortest of the three cases.


6.61<br />

subcritical flow<br />

flow <strong>pe</strong>r unit length q =Q /b<br />

b<br />

Total Energy Line<br />

2<br />

(V /2g)<br />

3<br />

2<br />

Y = Q /b<br />

0<br />

5 Y 0<br />

Y 0<br />

flow Q<br />

Ventilation air<br />

Sample of <strong>slurry</strong><br />

FIGURE 6-25 Sampling tailings with a moving bucket crossing the nap<strong>pe</strong> in a tailings drop box.<br />

travel of sampling bucket


D d<br />

D d<br />

Fig 6-30<br />

y c<br />

ventilation<br />

ventilation<br />

subcritical flow<br />

d f<br />

subcritical flow<br />

d f<br />

C<br />

L p<br />

L p<br />

B<br />

L d<br />

L<br />

j<br />

A<br />

L L<br />

d<br />

r<br />

L j<br />

d 1<br />

6.62<br />

d 3<br />

L r<br />

L d < L j Lr < L b<br />

Case (A)<br />

d<br />

1<br />

L < L<br />

L > L<br />

d j r b<br />

d < d 2<br />

d < d 3<br />

FIGURE 6-27 Patterns of flow with free fall with entry in a subcritical regime. (After Rand,<br />

1955.)<br />

d<br />

3<br />

FIGURE 6-26 Geometry of the nap<strong>pe</strong> from a waterfall.<br />

d<br />

d


D d<br />

SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

In Case C (Figure 6-28), the nap<strong>pe</strong> is submerged, and the depth of the tail water is<br />

higher than d 2. Compared with the previous two cases, this turbulent under-nap<strong>pe</strong> region<br />

is the dee<strong>pe</strong>st of the three cases.<br />

The turbulent roller extends much further and is less intense than the hydraulic jump.<br />

Referring to Figure 6-29, if D d is the depth of the fall from the bottom to the brink of the<br />

bottom of the drop box, a drop number Dr can be defined as<br />

2 q b<br />

�3 gD d<br />

6.63<br />

Dr = (6-89)<br />

In real life there is always a sort of churning area of liquid under the nap<strong>pe</strong>, but from a<br />

theoretical point of view, which ignores this pool of liquid, the location of the centerline of<br />

the nap<strong>pe</strong> intersecting with the bottom of the apron or the drop box would be expressed as<br />

= 1.98[Dr1/3 + 0.357 Dr2/3 ] 1/2 Lp � (6-90)<br />

Dd<br />

This point is also called the toe of the nap<strong>pe</strong>.<br />

The hydraulic jump occurs at a distance Ld, which may be smaller, equal to, or even<br />

larger than Lp (Figure 6-26) The ratio of Ld/Lp is maximum at 1.87 when Dr = 1.<br />

Tests reported by Rand (1955) indicate that the value of<br />

= 4.30Dr0.27 Ld � (6-91)<br />

Dd<br />

In cases where the hydraulic jump starts at the toe of the nap<strong>pe</strong>, the ex<strong>pe</strong>rimental work<br />

of Rand (1955) indicates that the reference depth d2 for the tail water can be expressed as<br />

a constant:<br />

Ld � = 2.60 (6-92)<br />

d2<br />

ventilation<br />

df<br />

L d<br />

d 1<br />

FIGURE 6-28 Geometry of nap<strong>pe</strong> from a free fall with entry in a critical regime. (After<br />

Rand, 1955.)<br />

L r<br />

L d = L j Lr = L b<br />

Case (B)<br />

d = d 2<br />

d 2


D d<br />

6.64 CHAPTER SIX<br />

ventilation<br />

d f<br />

Under the nap<strong>pe</strong>, a region of still water develops to a depth d f. The intersection of this<br />

rotating water with the nap<strong>pe</strong> is at point B of Figure 6-29. The height is d f, expressed as<br />

The height of the liquid d 1 is expressed as<br />

= Dr 0.22 (6-93)<br />

= 0.54Dr 0.425 (6-94)<br />

The height of the liquid d 2 in case (b) for entry in a critical regime is expressed as<br />

= 1.66Dr 0.27 (6-95)<br />

And the length to the intersection can be expressed by length L p or<br />

L pB = 1.98[Y 0(D d + 0.357Y 0 – d f)] 1/2 (6-96)<br />

The drop length or the length between the drop wall and the location of minimum<br />

depth of the liquid at the jump d j in Figure 6-26 at point A is expressed as<br />

L d<br />

� Dd<br />

L j<br />

d 1<br />

� Dd<br />

d 2<br />

� Dd<br />

1.98(1 + 0.357 Y0/Dd)�(Y� 0/ �D� �d)<br />

= ����<br />

(6-97)<br />

�[1� +� 0�.3�5�7�(Y� �D� 0/ d) � –� (�d� f/D� �]� d)<br />

Finally, the total length of the hydraulic jump from the point d j to the point where the<br />

tail–water has stabilized can be expressed as<br />

L r<br />

� Dd<br />

d<br />

Ld Lr d > d2 Case (C)<br />

d f<br />

� Dd<br />

L r > L b<br />

FIGURE 6-29 Free fall with a submerged nap<strong>pe</strong> (after Rand, 1955).<br />

d2 d1 = 6� � – �� (6-98)<br />

D Dd<br />

d


SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

These equations are based on pro<strong>pe</strong>r ventilation of the nap<strong>pe</strong>. If the nap<strong>pe</strong> is not pro<strong>pe</strong>rly<br />

ventilated, it becomes semiattached or totally attached to the drop box wall. This<br />

leads to a condition where flows may cause vibration of the drop box, which may tear it<br />

apart if it is not structurally designed to handle the vibration.<br />

The equations of Walter Rand were develo<strong>pe</strong>d for waterfalls. They are a good reference<br />

for designing drop boxes. Unfortunately, very little has been published over the<br />

years to examine the effect of solids on the level of turbulence at the toe of the nap<strong>pe</strong> and<br />

on the magnitude of the various parameters.<br />

Example 6-12<br />

A mass of liquid approaches a free fall at a Froude number of 1.0. The height of the liquid<br />

at the brink is measured to be 1.2 m (3.94 ft). The fall is 6 m (19.48 ft) deep. It is assumed<br />

that the width of the channel and drop box remain uniform. Determine the geometry of<br />

the hydraulic jump at the apron.<br />

Solution in SI Units<br />

From Equation 6-86:<br />

Yb � = 0.716<br />

Y0<br />

or Y0 = 1.2/0.716 = 1.676 m (or 5.499 ft).<br />

The Froude number of 1.0 occurs five times the critical depth upstream of the brink.<br />

2 1/3 The critical depth is defined as Y0 = [q b/g] , so<br />

qb = �(1�.6�7�6� 3 �·�9�.8�1�)�=� 6�.8�1� m� 2 �s� /<br />

From Equation 6-88, the drop number Dr is<br />

q b 2<br />

6.81 2<br />

6.65<br />

Dr = = = 0.0219<br />

The toe of the nap<strong>pe</strong> is determined from Equation 6-90:<br />

= 1.98 [Dr1/3 + 0.357 Dr2/3 ] 1/2 = 1.98 [0.02191/3 + 0.357 (0.02192/3 )] 1/2 = 1.098<br />

Lp = 1.098 × 6 = 6.6 m<br />

This point is also called the toe of the nap<strong>pe</strong>. The location of the hydraulic jump is obtained<br />

from Equation 6-91:<br />

= 4.30Dr0.27 = 4.3 × 0.02190.27 = 1.533<br />

Ld = 1.533 × 6 = 9.195 m<br />

The hydraulic jump occurs after the toe of the nap<strong>pe</strong>. Under the nap<strong>pe</strong>, a region of still<br />

water develops to a depth df, expressed by Equation 6-93 as<br />

= Dr0.22 = 0.02190.22 � ��<br />

3 3 (gDd) (9.81 · 6 )<br />

Lp �<br />

Dd<br />

Ld �<br />

Dd<br />

df � = 0.4314<br />

Dd<br />

df = 0.4314 · 6 = 2.59 m<br />

If this were <strong>slurry</strong>, it would be recommended to line this area to a height of 3 m by the<br />

length of Lp (6.59 m).<br />

The height of the liquid d1 is expressed by Equation 6-94:


6.66 CHAPTER SIX<br />

d 1<br />

� Dd<br />

= 0.54Dr 0.425 = 0.54 × 0.0219 0.425 = 0.1064<br />

d1 = 0.1064 × 6 = 0.6386 m<br />

The height of the liquid d2 is expressed by Equation 6-95:<br />

d 2<br />

� Dd<br />

= 1.66 Dr 0.27 = 1.66 × 0.0219 0.27 = 0.5916<br />

d2 = 0.5916 × 6 = 3.55 m<br />

The distance between d1 and d2 or length of the hydraulic jump is<br />

= 6 (d2/Dd – d1/Dd) = 6(0.5916 – 0.1064) = 2.911<br />

Lb = 2.911 × 6 = 17.47 m<br />

This length should be lined to the height of d2 + 10% or approximately 4 m.<br />

Solution in USCS Units<br />

From Equation 6-86:<br />

= 0.716<br />

or Y0 = 3.94/0.716 = 5.499 ft.<br />

The Froude number of 1.0 occurs five times the critical depth upstream from the brink.<br />

2 1/3 The critical depth is defined as Y0 = [qb/g] , so<br />

qb = (5.4993 · 32.2) = 73.17 ft2 Lr �<br />

Dd<br />

Yb �<br />

Y0<br />

/sec<br />

From equation 6-89, the drop number Dr is<br />

q b 2<br />

Dr = = 73.172 /(32.2 · 19.483 ) = 0.022<br />

The toe of the nap<strong>pe</strong> is determined from Equation 6-90:<br />

= 1.98[Dr1/3 + 0.357 Dr2/3 ] 1/2 = 1.98[0.0221/3 + 0.357 (0.0222/3 )] 1/2 = 1.099<br />

Lp = 1.099 × 19.48 = 21.4 ft<br />

This point is also called the toe of the nap<strong>pe</strong>. The location of the hydraulic jump is obtained<br />

from Equation 6-91:<br />

= 4.30Dr0.27 = 4.3 × 0.0220.27 � 3 (gDd) Lp �<br />

Dd<br />

Ld � = 1.53<br />

Dd<br />

Ld = 1.53 × 19.48 = 29.80 ft<br />

The hydraulic jump occurs after the toe of the nap<strong>pe</strong>. Under the nap<strong>pe</strong>, a region of still<br />

water develops to a depth df, expressed by Equation 6-93 as<br />

d f<br />

� Dd<br />

= Dr 0.22 = 0.022 0.22 = 0.432


SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

df = 0.432 · 19.48 = 8.42 ft<br />

If this were <strong>slurry</strong>, it would be recommended to line this area to a height of 10 ft by the<br />

length of Lp or approximately 21.6 ft.<br />

The height of the liquid d1 is computed from Equation 6-94:<br />

d 1<br />

� Dd<br />

= 0.54Dr 0.425 = 0.54 × 0.022 0.425 = 0.1064<br />

d1 = 0.1064 × 19.48 = 2.07 ft<br />

The height of the liquid d2 is computed from Equation 6-95:<br />

d 2<br />

� Dd<br />

= 1.66 Dr 0.27 = 1.66 × 0.022 0.27 = 0.5916<br />

d2 = 0.5916 × 19.42 = 11.49 ft<br />

The distance between d1 and d2 or length of the hydraulic jump is<br />

Lb � = 6 (d2/Dd – d1/Dd) = 6(0.5916 – 0.1064) = 2.911<br />

Dd<br />

Lb = 2.911 × 19.48 = 56.71 ft<br />

This length should be lined to the height of d2 + 10% or approximately 12.6 ft.<br />

6-12 PLUNGE POOLS AND DROPS<br />

FOLLOWED BY WEIRS<br />

In nature, the scouring depth of a waterfall may be typically one third of the depth of the<br />

waterfall. An example of an engineering exercise along these lines was the construction<br />

of Mossyrock spillway on the Colwitz River near Tacoma, Washington (U.S.A.). The<br />

spillway was created to handle a 183 m (600 ft) drop.<br />

In the case of slurries, the wear is accelerated by the very nature of the abrasive and<br />

erosive particles. S<strong>pe</strong>nt mill liners, s<strong>pe</strong>nt mill balls, steel grading, and s<strong>pe</strong>nt pump liners<br />

are installed at the bottom of drop boxes to prevent wear. It is not always cost effective to<br />

design for a scouring depth equal to one third of the free fall.<br />

A drop box can be ex<strong>pe</strong>nsive to construct. One of the largest <strong>slurry</strong> drop boxes was<br />

built by Fluor Daniel for the Caujone mine owned by the Southern Peru Cop<strong>pe</strong>r Corporation<br />

in Peru. It was designed to handle a tailing flow of 7.3 m 3 /s (116,000 gpm). The drop<br />

was 10 m (32 ft) (Figures 6-24 and 6-30) deep and the <strong>slurry</strong> had to be redirected under an<br />

existing truck road. The author was the hydraulic engineer on the project.<br />

To reduce the length of the pond, it is recommended to add a weir (Windsor, 1938).<br />

This alternative method is included in the discussion of the pa<strong>pe</strong>r of Moore (1943) by L.<br />

S. Hall (1943). On the basis of the work of Blackhmereff (1936), Hall develo<strong>pe</strong>d an approach<br />

to reduce the length of the transition region at the toe of the nap<strong>pe</strong> by adding a<br />

weir. The weir raises the water level and causes the nap<strong>pe</strong> to impinge water at a higher<br />

point of intersection.<br />

Referring to Figure 6-30, the length of the pond can be reduced to L�. If D d is the depth<br />

of the drop, an energy line E 0 is defined as<br />

E 0 = D d + 1.5Y 0<br />

6.67<br />

(6-99)


6.68 CHAPTER SIX<br />

E 0<br />

Fig 6 - 32<br />

Y<br />

0<br />

/2<br />

D d<br />

D d<br />

Y 0<br />

The level of the liquid over the weir Z 0 can be expressed graphically as in Figure 6-32<br />

or mathematically as in the following equation:<br />

(Y0/d1) d1 = + – 1.5 (6-100)<br />

2<br />

�2 2�<br />

(Y0/d1) = – � + 1.0 (6-101)<br />

2d1<br />

3<br />

�2 2�<br />

= 1 + – � –1 + ��1� +� 1� 6� �� 2��� +��� –� 1������<br />

where � is determined from the following cubic equation:<br />

(6-102)<br />

– �2 � + 3� + 2�2 Z0 3Y0 d1 Dd 3Y0 � � � � �<br />

Dd 2Dd 2Dd<br />

d1 2d1<br />

Y0 2Dd = 0 (6-103)<br />

Y 0 3<br />

� d 1 3<br />

D d<br />

� Y0<br />

D d<br />

� d1<br />

� d1<br />

� Y0<br />

De<strong>pe</strong>nding on the amount of energy dissipation before the location of d 1, � may be assumed<br />

to be 1.0 for no dissipation at all (Bakhemeteff, 1932) or as low as 0.95 for some<br />

dissipation before the jump (Bobin, 1934):<br />

3Y 0<br />

steep cascade at inlet<br />

Z0 = Dd + � – d2 – hw (6-104)<br />

2<br />

where hw is the height of the weir that controls the plunge pool relative to the apron.<br />

The length of the plunge pool is expressed as:<br />

Y0 L� = C���� �<br />

Dd<br />

� Y0<br />

3Y 0<br />

1� +���� Y� 0D� d��<br />

(6-105)<br />

d 2<br />

Z 0<br />

h w<br />

L' L'<br />

2 L'<br />

FIGURE 6-30 A weir to control the flow of <strong>slurry</strong> from the nap<strong>pe</strong> of a drop box. (After Hall,<br />

1943 in his discussion of Moore, 1943.)


Z / D<br />

0 d<br />

1.0<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

where C can equal 1.7 for low spray but can also equal as high as 2.0 for significant spray.<br />

Standish Hall (1943) proposed that length L� be followed by an equal transition.<br />

Example 6-13<br />

Referring to Example 6-12, determine the length of the plunge pool if a controlling weir<br />

is added. Determine the level of the liquid Z 0.<br />

Solution in SI Units<br />

The critical depth was determined to be 1.676 m. The drop is 6 m. Assuming C = 2.0,<br />

Referring to Figure 6-25:<br />

SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

L� = 2�[1�.6�7�6� ·� 6� +� 1�.6�7�6� 2 � ] = 7.17 m<br />

Z 0<br />

� Dd<br />

Z / D<br />

0 d<br />

Y 0<br />

� Dd<br />

d / D<br />

1 d<br />

=<br />

1.676<br />

��<br />

6 = 0.279<br />

� 0.84 or Z 0 � 0.84 × 6 = 5.04 m<br />

2.0<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

6.69<br />

0.0<br />

0.0<br />

0.4 0.8 1.2<br />

Y / D<br />

0 d<br />

1.6 2.0<br />

FIGURE 6-31 Curves to determine the height of the weir in a plunge pool.(After Hall, 1943<br />

in his discussion of Moore, 1943, by <strong>pe</strong>rmission of ASCE.)<br />

Since Z0 is measured from E0, and<br />

E0 = Dd + 1.5 Y0 = 6 + 1.5 × 1.676 = 8.51 m<br />

the liquid level is 8.51 – 5.04 = 3.47 m above the apron.<br />

If the engineer builds a weir 2 m high (hw) it will be submerged by a depth of 1.47 m,<br />

corresponding to the value of d2. 0.6<br />

0.4<br />

0.2<br />

d / D<br />

1 d


6.70 CHAPTER SIX<br />

Solution in USCS Units<br />

The critical depth was determined to be 5.5 ft. The drop is 19.48 ft. Assuming C = 2.0,<br />

Referring to Figure 6-25,<br />

FIGURE 6-32 Walls of a weir showing sediment coating.<br />

Z 0<br />

� Dd<br />

L� = 2�[5�.5� ·� 1�9�.4�8� +� 5�.5� 2 � ] = 23.44 ft<br />

Y 0<br />

� Dd<br />

5.5<br />

= � = 0.279<br />

19.48<br />

� 0.84 or Z 0 � 0.84 × 19.48 = 16.36 ft<br />

Since Z0 is measured from E0, and<br />

E0 = Dd + 1.5 Y0 = 19.48 + 1.5 × 5.5 = 27.73 ft<br />

the liquid level is 27.73 – 16.36 = 11.1 ft above the apron.<br />

If the engineer builds a weir 6.56 ft high (hw) it will be submerged by a depth of 4.82<br />

ft, corresponding to the value of d2. The flow of <strong>slurry</strong> in flumes and through drop boxes is fairly complex and under certain<br />

conditions hydraulic jumps occur with considerable turbulence. For fairly abrasive<br />

slurries, wear is a concern. In other situations such as cop<strong>pe</strong>r mines, the presence of lime<br />

in the <strong>slurry</strong> may actually end up coating the flume with deposited lime that consolidates<br />

with time. This deposition of lime or similar sediments coats the flume, but does completely<br />

change the roughness of the wall (Figure 6-33). In some cases the designer must<br />

try to avoid break up the transported solids such as coal (Kuhn, 1980).


Values of y/y a<br />

-3.0<br />

+1.0<br />

+0.5<br />

0.0<br />

-0.5<br />

-1.0<br />

-1.5<br />

-2.0<br />

S<strong>pe</strong>cial transition areas may be lined with abrasion resistant steel or with rubber. The<br />

rubber is glued to steel plates that are bolted to the concrete (see Figure 6-1).<br />

The analyses of Hall (1943) and Moore (1941,1943) are based on the assumption that<br />

the liquid enters the fall from a subcritical regime, with minimum energy, and accelerates<br />

at the brink. The projection of the nap<strong>pe</strong> and contact with the apron is even more complicated<br />

when the jet approaches the brink at su<strong>pe</strong>rcritical flows. Rouse, in his discussion of<br />

Moore (1943), discussed the changes in Froude numbers of 1–14 (Figure 6-30).<br />

6-13 CONCLUSION<br />

SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

Slurry flows in o<strong>pe</strong>n channels are fairly complex but they follow many of the principles<br />

of closed conduit flows discussed in the previous two chapters. When the s<strong>pe</strong>ed is insufficient<br />

or the Froude number is low, deposition occurs and dunes or a stationary bed form.<br />

Since most books on <strong>slurry</strong> flows are focused on pi<strong>pe</strong> flows, this chapter presented an<br />

exhaustive review of the mathematics of o<strong>pe</strong>n channel <strong>slurry</strong> flows and design of drop<br />

boxes. The practical engineer should find in the worked examples a methodology to apply<br />

such complex equations. It is ho<strong>pe</strong>d that new generations of academicians and students<br />

will enrich the understanding of such complex flows. The design of o<strong>pe</strong>n channel flows<br />

requires frequent iterations for slo<strong>pe</strong>, stability (Froude number), roughness, etc. The use<br />

of modern <strong>pe</strong>rsonal computers with the appropriate equations allows the engineer to optimize<br />

the hydraulic design.<br />

On a note of caution, the design engineer should not apply data from small to large<br />

flumes. The change of the hydraulic radius and the ratio of particle size to depth of flow<br />

affect the magnitude of the slo<strong>pe</strong> of the launder.<br />

NOMENCLATURE<br />

-2.0 -1.0 0.0 1.0 2.0 3.0 4.0<br />

-2.0 -1.0 0.0 1.0 2.0 4.0<br />

Values of x/y a<br />

Fr = 2.18<br />

Fr = 1.8<br />

Fr = 1<br />

a Nondimensional parameter and function of Hedstrom number<br />

a Reference depth for concentration calculations<br />

Fr = 3.02<br />

Fr=4.12<br />

6.71<br />

FIGURE 6-33 Effect of the Froude number at the entry to the waterfall on the sha<strong>pe</strong> of the<br />

nap<strong>pe</strong>. [After Rouse (1943) in his discussion of Moore (1943).]<br />

6-14


6.72 CHAPTER SIX<br />

Ab Area of the horizontal projection of the lee face of the bed forms<br />

b Nondimensional parameter<br />

bw Wetted width<br />

C Time-averaged concentration of sus<strong>pe</strong>nded solids<br />

Ca Concentration at height “a”<br />

CD Drag coefficient of particles for a heterogeneous <strong>slurry</strong><br />

Ch Chezy number<br />

CL Lift coefficient<br />

Cm Depth-averaged concentration of solids<br />

CT Mean transport concentration of solid particles in the <strong>slurry</strong> mixture<br />

Cv Volume fraction of solid particles in the <strong>slurry</strong> mixture<br />

Cw Weight fraction of solid particles in the <strong>slurry</strong> mixture<br />

Cy Volume fraction of solid particles in the <strong>slurry</strong> mixture at level “y”<br />

d Depth<br />

db Depth at which a stationary hydraulic jump occurs between the su<strong>pe</strong>rcritical and<br />

subcritical flows on the apron after a free fall<br />

df Depth of under nap<strong>pe</strong> liquid between drop wall and nap<strong>pe</strong><br />

dj Depth at the hydraulic jump on the apron from a free fall<br />

dp Diameter of the particle<br />

dt Final depth of the tail water after the hydraulic jump due to fall<br />

d1 Depth at the toe of the nap<strong>pe</strong> for a free fall and drop<br />

d2 Reference depth for subcritical tail water after the free fall in the case of a hydraulic<br />

jump occurring at the toe of the nap<strong>pe</strong><br />

d3 Depth of su<strong>pe</strong>rcritical flow at beginning of the hydraulic jump downstream of the<br />

nap<strong>pe</strong><br />

d50 Particle diameter passing 50% (m)<br />

d85 Particle diameter passing 85% (m)<br />

Dd Depth of drop box of free-fall drop<br />

DH Hydraulic diameter<br />

DI Conduit inner diameter (m)<br />

Dr Drop number for free fall<br />

Er Coefficient correlating relative roughness to friction and average velocity<br />

E0 Total energy level for a free-fall problem of a liquid relative to the apron<br />

fD Darcy friction factor<br />

fD� Darcy friction factor for the channel without bed forms<br />

fD�� Darcy friction factor due to the bed forms<br />

fDL Darcy friction factor for liquid<br />

fN Fanning friction factor<br />

f1 Mathematical function<br />

f2 Mathematical function<br />

fNL Laminar component of fanning friction factor<br />

FN fluid force normal to the direction of flow<br />

Fr Froude number<br />

fT Turbulent component of fanning friction factor<br />

FT Fluid force tangent to the direction of flow<br />

g Acceleration due to gravity (9.81 m/s2 )<br />

G Flocculation gradient<br />

h Head due to friction losses<br />

ha Depth ratio defined by Equation 6-31<br />

hw Height of weir in a plunge pool with a weir<br />

He Hedstrom number


SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

6.73<br />

J Nondimensional parameter to account for dynamic viscosity in deposit velocity<br />

ks Linear roughness (m)<br />

Ke Ex<strong>pe</strong>rimental constant<br />

Kx Von Karman coefficient<br />

L Length of conduit<br />

L� Length of drop pool with a controlling weir<br />

Lb Distance between the point of impingement of the nap<strong>pe</strong> and the tail water depth<br />

Ld Distance between drop wall and toe of the nap<strong>pe</strong> for a free-fall drop<br />

Lj Distance between the wall of the free fall and the hydraulic jump on the apron<br />

Lmix Mixing length for eddies<br />

Lp Theoretical distance to intersection of the center of the nap<strong>pe</strong> and the bottom of the<br />

drop box with under-nap<strong>pe</strong> pool (see Figure 6-17)<br />

Lr Total length to the stable tail water<br />

m Exponent from the Darby equation<br />

M Montuori number<br />

n Manning roughness number<br />

qb Flow rate <strong>pe</strong>r unit width of launder (m2 /s)<br />

qbs Flow rate of sediments <strong>pe</strong>r unit width<br />

Q Flow rate (m3 /s)<br />

P Power<br />

Patm Atmospheric pressure<br />

PL Plasticity number<br />

Pw Wetted <strong>pe</strong>rimeter<br />

R Radius<br />

Re Reynolds number<br />

Rep Particle Reynolds number<br />

RH Hydraulic radius (m)<br />

RH� Hydraulic radius due to grain roughness<br />

RH�� Hydraulic radius due to bedforms<br />

S Slo<strong>pe</strong><br />

Sm S<strong>pe</strong>cific gravity of mixture<br />

U Horizontal component of velocity<br />

U� Horizontal component of velocity due to turbulence<br />

Uav Average s<strong>pe</strong>ed<br />

Ub Bed velocity<br />

Ubc Critical velocity to start the motion of the bed<br />

Ucr Critical velocity to start the flow of cohesive elements<br />

Uf friction velocity<br />

Uf� Friction velocity due grain roughness<br />

Uf�� Friction velocity due to dunes or bedforms<br />

Um Average s<strong>pe</strong>ed<br />

Umax Maximum s<strong>pe</strong>ed<br />

V Average velocity of flow (m/s)<br />

V� Average vertical velocity due to eddies<br />

VC Camp minimum self-cleaning velocity for a sewer (m/s)<br />

VD Deposit velocity in a launder (m/s)<br />

Ve Verdinokov number<br />

Vm Mean vertical velocity component<br />

Vsc Self-cleaning velocity of a launder<br />

Vt Particle terminal velocity<br />

Vol Volume


6.74 CHAPTER SIX<br />

w Width of launder<br />

x Local horizontal ordinate<br />

X0 A coefficient of cohesion of the material<br />

y Local vertical coordinate in the launder<br />

ym Average depth of the <strong>slurry</strong> in the launder<br />

Y Depth of launder<br />

Y0 Critical depth of the liquid at Froude number of one<br />

Z Function of the height above the bed of a launder<br />

Z0 Depth of liquid surface in a plunge pool over the weir<br />

Z1 Empirical function of grain distribution above bed<br />

Greek letters<br />

� Angle of inclination of flow with res<strong>pe</strong>ct to particle<br />

� Constant of proportionality<br />

� Constant of proportionality in Celik’s equation<br />

�m Coefficient of exchange of momentum between neighboring streams of the fluid<br />

�s Mass transfer coefficient<br />

� Angle of slo<strong>pe</strong><br />

� Factor of energy dissipation before the hydraulic jump in a free fall<br />

� A<br />

Graf–Acaroglu function<br />

� Coefficient of rigidity<br />

� Data about cohesion<br />

� tan –1 S<br />

� Wavelength of deposited dunes and antidunes<br />

� Absolute (or dynamic) viscosity<br />

�m Absolute (or dynamic) viscosity of mixture<br />

� Dynamic viscosity<br />

� Shear stress<br />

�cr Critical shear stress<br />

�L Fluid shear stress<br />

�0 Yield stress for Bingham plastics and pseudoplastics<br />

�w Shear stress at the wall<br />

� Density<br />

�L Density of carrier liquid<br />

�m Density of <strong>slurry</strong> mixture (Kg/m3 )<br />

�s Density of solids in mixture (Kg/m3 )<br />

� Exponent for effective shear stress � 0.06<br />

� Sedimentation coefficient<br />

�A Graf–Acaroglu function<br />

�D Sha<strong>pe</strong> factor<br />

�1 Sha<strong>pe</strong> factor<br />

�2 Sha<strong>pe</strong> factor<br />

�3 Sha<strong>pe</strong> factor<br />

6-15 REFERENCES<br />

Abulnaga, B. E. 1997. Channel 1.0 Computer Program for O<strong>pe</strong>n Channel Slurry Flows. Develo<strong>pe</strong>d<br />

for Fluor Daniel Wright Engineers. Internal report.<br />

Acaroglu, E. R. 1968. Sediment Transport in Conveyance Systems. Ph.D. diss., Cornell University.


SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

6.75<br />

Ambrose H. H. 1953. The transportation of sand in pi<strong>pe</strong>s with free surface flow. In Proceedings of<br />

the Fifth Hydraulics Conference. Ames: State University of Iowa, pp. 77–88.<br />

The American Society of Civil Engineers. 1975. Sedimentation Engineering. Manuals and Reports<br />

on Engineering Practice. No. 54. New York: ASCE.<br />

The American Society of Civil Engineers and the Water Pollution Control Federation. 1977. Wastewater<br />

Treatment Plant Design. ASCE Manual and Reports on Engineering Practice No. 36.<br />

(Also published as WCF Manual of Practice No. 8.)<br />

Apmann, R. P., and R. R. Rumer, Jr. 1967. Diffusion of Sediments in a Non-Uniform Flow Field. Report<br />

prepared for the Department of Civil Engineering, Faculty of Engineering and Applied<br />

Science, State University of New York at Buffalo. Report No. 16.<br />

Bakhmeteff, B. A. 1932. Hydraulics of O<strong>pe</strong>n Channels. New York: McGraw-Hill.<br />

Blench, T., V. J. Galay, and A. W. Peterson. 1980. Steady fluid-solid flow in flumes. Pa<strong>pe</strong>r C-1, presented<br />

at the 7th Annual Hydrotransport Conference, Sendai, Japan. BHR Group.<br />

Bobin, P. M. 1934. The design of stilling basins. Transactions of the Scientific Research Institute of<br />

Hydrotechnics, XIII, 79–123.<br />

Bogardi, J. L. 1965. Euro<strong>pe</strong>an concepts of sediment transportation. Proc. Am. Soc. Civil Engineers,<br />

91, HY1, 29–54.<br />

Boussinesq, M. J. 1877. (Ed.). Essai sur la Theorie des Eaux Courantes. [A Study on the Theory of<br />

Flowing Waters.] Memoires, Presentèes par Divers Savants—L’Academie de l’Institut de<br />

France, 23, 1–680. [Transactions of the French Academy Institute, 23, 1–680.]<br />

Brush, L. M., H. W. Ho, and S. R. Singamsetti. 1962. A study of sediment in sus<strong>pe</strong>nsion. Intern. Assoc.<br />

Sci. Hydr., Commiss. Land Erosion, No. 59.<br />

Camp, T. R. 1955. Flocculation and flocculation basins. Transactions Am. Soc. of Civil Engineers,<br />

120, 1 1–16.<br />

Celik, I., and W. Rodi. 1984. A Deposition-Entrainment Model for Sus<strong>pe</strong>nded Sediment Transport.<br />

Internal Report prepared by the University of Karlsruhe, Germany. Report No.<br />

SFB210/T/6.<br />

Celik, I., and W. Rodi. 1991. Sus<strong>pe</strong>nded sediment-transport capacity for o<strong>pe</strong>n channels. Journal of<br />

Hydraulic Engineering, 117, 2, 191–204.<br />

Chien, N. 1954. The present status of research on sediment transport. Proc. Am. Soc. Civil Engrs.,<br />

80, No 565, 33.<br />

Coo<strong>pe</strong>r, R. H. 1970. A study of bed Material Transport Based on the Analysis of Flume Ex<strong>pe</strong>riments.<br />

PhD. thesis, Department of Civil Engineering, University of Alberta, Canada.<br />

Dominguez, B., R. Souyris, and A. Nazer. 1996. Deposit velocity of <strong>slurry</strong> flow in o<strong>pe</strong>n channels.<br />

Pa<strong>pe</strong>r read at the symposium, Slurry Handling and Pi<strong>pe</strong>line Transport. Thirteenth annual International<br />

Conference of the British Hydromechanic Research Association, Johannesburg, South<br />

Africa.<br />

Einstein H. A. 1950. The Bed-Load Function for Sediment Transportation in O<strong>pe</strong>n Channel Flows.<br />

Technical Bulletin No. 1026. U.S. Deptartment of Agriculture Soil Conservation Service.<br />

Einstein H. A. and J. A. Hardner, 1954. Velocity distribution and boundary layer at channel bends.<br />

Am. Geophysical Union Trans., 35, 114–120.<br />

Einstein, H. A., and N. Chien. 1955. Effects of Heavy Sediment Concentration Near the Bed on Velocity<br />

and Sediment Distribution. MRD Sed. Ser. Berkeley: University of California.<br />

Fathy, A., and M. A. Shaarawi. 1954. Hydraulics of free overfall. Proc. Am. Soc. Civ. Eng, 80, 564,<br />

1–12.<br />

Fortier, S., and F. C. Scobey. 1925. Permissible canal velocities. Trans. Am. Soc. Civil Engrs, 51, 7,<br />

1397–1413.<br />

Garde, R. J., and J. Dattari. 1963. Investigation of the total sediment load of streams. Res. J. University<br />

of Roorkee. Internal report.<br />

Graf, W. H. 1971. Hydraulics of Sediment Transport. New York: McGraw-Hill.<br />

Graf, W. H., and E. R. Acaroglu. 1968. Sediment transport in conveyance <strong>systems</strong>. Part I. Bulletin.<br />

Intern. Association of Sci. Hydr., 2.<br />

Green, H. R., D. H. Lamb, and A. D. Tylor. 1978. A new launder design procedure. Pa<strong>pe</strong>r read at the<br />

Annual Meeting of the Society of Mining Engineers, March, Denver, Colorado.<br />

Grim, R. E. 1962. Applied Clay Mineralogy. New York: McGraw-Hill.<br />

Guy, H. P., R. E. Rathbun, and E. V. Richardson. 1967. Recirculation and sand-feed flume ex<strong>pe</strong>riments.<br />

Pa<strong>pe</strong>r 5428. Am. Soc. of Civil Eng., 93 HYS, 97–114, Sept.


6.76 CHAPTER SIX<br />

Hall, S. L. 1943. Discussion to pa<strong>pe</strong>r by W. L. Moore. 1943. Energy loss at the base of the free overfall.<br />

Transaction of the A.S.C.E., 108, 1378–1387.<br />

Henderson, F. M. 1990. O<strong>pe</strong>n Channel Flow. New York: Macmillan.<br />

Ismail, H. M. 1952. Turbulent transfer mechanism and sus<strong>pe</strong>nded sediments in closed channels.<br />

Trans. ASCE, 117, 409–447.<br />

Julian, Smart and Allan. 1921. Cyaniding Gold and Silver Ores. Internal report presented to J. B.<br />

Lip<strong>pe</strong>nicott Co., U.S.A. Reported by Tournier and Judd (1945).<br />

Karasev, I. F. 1964. The regimes of eroding channels in cohesive materials. Soviet Hydrol. (Am.<br />

Geophysics Union), Vol. 6.<br />

Kennedy, J. F. 1963. The mechanics of dunes and antidunes in erodible bed channels. Journal Fluid<br />

Mech., 16, 4.<br />

Keulegan, G. H. 1938. Laws of turbulent flow in o<strong>pe</strong>n channels. Journal of Research (National Bureau<br />

of Standards, U.S. Dept of Commerce), 21, 707–741.<br />

Kuhn, M. 1980. Hydraulic Transport of solids in flumes in the mining industry. Pa<strong>pe</strong>r C3 read at the<br />

7th International Conference of the Hydraulic Transport of Solids in Pi<strong>pe</strong>s, Sendai, Japan.<br />

Cranfield, UK: BHRA Fluid Engineering, pp. 111–122.<br />

Liu, H. K. 1957. Mechanics of sediment—Ripple formation. Proc. Am. Soc. Civil. Eng., 83, HY2,<br />

Pa<strong>pe</strong>r 1197.<br />

Lovera, F., and J. F. Kennedy. 1969. Friction factor for flat-bed flows in sand channels. Proc. Am.<br />

Soc. Civil Eng., 95, HY4, Pa<strong>pe</strong>r 6678, pp. 1227–1234.<br />

Majumdar, H., and M. R. Carstens. 1967. Diffusion of Particles by Turbulence: Effect of Particle<br />

Size. Water Res. Center, Report WRC-0967, Georgia Inst. Techn., Atlanta, U.S.A.<br />

Manning R.1895. On the flow of o<strong>pe</strong>n channels and pi<strong>pe</strong>s. Transactions, Institution of Civil Engineers<br />

of Ireland, 10, 14, 161–207.<br />

Matyukhin, V. J., and O. N. Prokofyev. 1966. Ex<strong>pe</strong>rimental determination of the coefficient of vertical<br />

turbulent diffusion in water for settling particles. Soviet Hydrol. (Am. Geophys.Union), No<br />

3.<br />

Ministry of Technology of the United Kingdom. 1969. Charts for the Hydraulic Design of Channels<br />

and Pi<strong>pe</strong>s. London: Ministry of Technology of the United Kingdom.<br />

Moore, W. L. 1943. Energy loss at the base of the free overfall. Transaction of the A.S.C.E., 108,<br />

1343–1392.<br />

Neil, C. R. 1967. Mean velocity criterion for scour of coarse uniform bed material. In International<br />

Association of Hydrology Research, 12th Congress. Fort Collins, CO.<br />

Nie<strong>pe</strong>lt, W. A., and F. A. Locher. 1989. Instability in high velocity <strong>slurry</strong> flows. Mining Engineering,<br />

41, 12, 1204–1209.<br />

O’Brien, M. P. 1933. Review of the theory of turbulent flow and its relation to sediment transportation.<br />

Trans. Am. Geophysics, 14, 487–491.<br />

Rand, W. 1955. Flow geometry at straight drop spillways. Transaction of the Am. Soc. Civ. Eng., 81,<br />

791, 1–13.<br />

Reynolds, O. 1895. On the Dynamical theory of incompressible viscous fluids and the determination<br />

of the criterion. Catalogue of Scientific Pa<strong>pe</strong>rs, compiled by the Royal Society of London, Vol.<br />

2, pp. 535–577. Cambridge, UK: Cambridge University Press.<br />

Richardson, E. G. 1937. The sus<strong>pe</strong>nsion of solids in a turbulent stream. Proceedings of the Royal Society<br />

of London, 162, Series A, 583–597.<br />

Richardson, E. V., and D. B. Simons. 1967. Resistance to flow in sand channels. Pa<strong>pe</strong>r read at International<br />

Association Hydrology Research, 12th Congress, Fort Collins, Colorado.<br />

Rouse, H. 1937. Modern conceptions of the mechanics of fluid turbulence. Transactions of the Am.<br />

Soc. Of Civil Engrs., 102, 536.<br />

Rubin, M. B. 1997. Relationship of critical flow in waterfall to minimum energy head. Journal of<br />

Hydraulics, 123, January, 82–84.<br />

Silberman, E. 1963. Friction factors in o<strong>pe</strong>n channels. Proc. Am. Soc. Civil Engrs., 89, no. HY2,<br />

Simons, D. B. and M. L. Albertson. 1963. Univorm water conveyance in alluvial channels. Proc.<br />

Am. Soc. Civ. Eng., 128, 1.<br />

Slatter, P. T., G. S. Thorvaldsen, and F. W. Petersen. 1996. Particle roughness turbulence. Pa<strong>pe</strong>r read<br />

at the 13th International Conference on Slurry Handling and Pi<strong>pe</strong>line Transport, at British Hydromechanic<br />

Research Association, Johannesburg, South Africa.<br />

Shook, C. A. 1981. Lead Agency Report II For Coarse Coal Transport. MTCH Hydrotransport Coo<strong>pe</strong>rative<br />

Programme. Saskatoon, Canada: Saskatchewan Research Council.


SLURRY FLOW IN OPEN CHANNELS AND DROP BOXES<br />

6.77<br />

Stricklen, R. 1984. Slurry handling considerations. Pa<strong>pe</strong>r read at the 1984 Annual Meeting of the<br />

American Institute of Mining Engineering, Denver, Colorado, U.S.A.<br />

Thomas A. D. 1979.The role of laminar/turbulent transition in determining the critical deposit velocity<br />

and the o<strong>pe</strong>rating pressure gradient for long distance <strong>slurry</strong> pi<strong>pe</strong>lines. Pa<strong>pe</strong>r read at the 6th<br />

International Conference of the Hydraulic Transport of Solids in Pi<strong>pe</strong>s. Cranfield, UK: BHRA<br />

Fluid Engineering, pp. 13–26.<br />

Tournier, E. J. and E. K. Judd. 1945. Storage and mill transport. In Handbook of Mineral Dressing—<br />

Ore and Industrial Minerals. New York: Wiley.<br />

Vanoni, V. A. 1946. Transportation of sus<strong>pe</strong>nded sediment by water. Pa<strong>pe</strong>r no. 2267 Trans. Am. Soc.<br />

Civ. Eng. Hydraulics Division, 111, 67–133.<br />

Vanoni, V. A., and L. S. Hwang. 1967. Relation between bedforms and friction in streams. Proc. Am.<br />

Soc. Civil. Engrs. 93, no. HY3,<br />

Van Rijn, L. C. 1981. Comparison of Bed-Load Concentration and Bed-Load Transport. Report prepared<br />

by the Delft Hydraulic Laboratory, Delft, The Netherlands. Report No. S 487, Part I.<br />

Von Karman, T. 1934. Turbulence and skin friction. Journal of Aeronautical Sciences, 1, 1, 1–20.<br />

Von Karman, T. 1935. Some as<strong>pe</strong>cts of the turbulence problem. Mechanical Engineering, 57,<br />

407–412.<br />

Wasp, E., J. Penny, and R. Ghandi. 1977. Solid-Liquid Flow Slurry Pi<strong>pe</strong>line Transportation. Aedermannsdorf,<br />

Switzerland: Trans Tech Publications.<br />

Whipple, K. X. 1997. O<strong>pe</strong>n channel flow of Bingham fluids. Journal of Geology, 105, 243–262.<br />

Wilson, K. C. 1991. Slurry transport in flumes. In Slurry Handling, edited by N. P. Brown and N. I.<br />

Heywood. New York: Elsevier Applied Sciences.<br />

Windsor, L. M. 1938. The barrier system of flood control. Civil Engineering (October), 675.<br />

Wood P.A. 1980. Optimization of flume geometry for o<strong>pe</strong>n channel transport . Pa<strong>pe</strong>r C2 read at the<br />

7th International Conference of the Hydraulic Transport of Solids in Pi<strong>pe</strong>s, Sendai , Japan.<br />

Cranfield, UK: BHRA Fluid Engineering, pp. 101–110.<br />

Yalin, M. S. 1977. Mechanics of Sediment Transport. 2nd Edition. Toronto: Pergamon Press.<br />

Zip<strong>pe</strong>, H. J., and H. Graf. 1983. Turbulent boundary-layer flow over <strong>pe</strong>rmeable and non-<strong>pe</strong>rmeable<br />

rough surfaces. J. Hydr. Res., 21, 1, 51–65.<br />

Further readings<br />

Bagnold, R. A. 1955. Some flume ex<strong>pe</strong>riments on large grains but little denser than the transporting<br />

fluid and their implication. Part 3. Proc. Inst. Civil Engrs, 4. 174–205.<br />

Gilbert, G. K. 1914. Transportation of Debris by Running Water. Pa<strong>pe</strong>r no. 86. U.S. Geological Survey.<br />

Guy, H. P., D. B. Simons, and E. V. Richardson. 1966. Summary of Alluvial Channel Data From<br />

Flume Ex<strong>pe</strong>riments, 1956–1961. Pa<strong>pe</strong>r No. 462-I. U.S. Geological Survey.<br />

Khurmi, R. S. 1970. Hydraulics and hydraulic machines. Delhi: S. Chand & Co.<br />

Lacey, G. 1930. Stable channels in alluvium. Pa<strong>pe</strong>r no. 4736. Proc. Inst. Civil Engs., 229, 529–384.<br />

Lacey, G. 1934. Uniform flow in alluvial rivers and canals. Pa<strong>pe</strong>r no. 237. Proc. Inst. Civil Engs.,<br />

237, 421–544.<br />

Lacey, G. 1947. A general theory of flow in alluvium. Pa<strong>pe</strong>r no. 5518. Journal Inst. Civil Engs., 17,<br />

1, 16–47.<br />

Nino, Y., and M. Garcia. 1998. Ex<strong>pe</strong>riments on saltation of sand in water. Journal of Hydraulics,<br />

124, 10, 1014–1025.<br />

Turton, R. K. 1966. Design of <strong>slurry</strong> distribution manifolds. Engineer, 221, 641–643.<br />

Wilson, K. C. 1980. Analysis of <strong>slurry</strong> flows with a free surface. Pa<strong>pe</strong>r C4 read at Hydrotransport 7,<br />

Sendai, Japan. Cranfield, UK: BHRA Group, pp 123–132.


CHAPTER 11<br />

SLURRY PIPELINES<br />

11-0 INTRODUCTION<br />

Engineering as a science must balance sophisticated mathematical models with practical<br />

field ex<strong>pe</strong>rience. In this chapter, some s<strong>pe</strong>cific cases of <strong>slurry</strong> pi<strong>pe</strong>lines will be examined<br />

for coal, iron sand, clay, phosphate, limestone, and other materials. The practical<br />

ex<strong>pe</strong>rience with these minerals in different forms, particles sizes, and volumetric concentrations<br />

is very useful for the design of new pi<strong>pe</strong>lines or the modification of existing<br />

<strong>systems</strong>.<br />

As in the case of coal, which can be pum<strong>pe</strong>d in very fine sizes—smaller than 100–125<br />

�m (mesh 140–120) and as coarse as 50 mm (2 in), all kinds of schemes and equipment<br />

have been develo<strong>pe</strong>d. The economics of preparing the <strong>slurry</strong> at the feed point of the<br />

pi<strong>pe</strong>line, and the cost of building and o<strong>pe</strong>rating the pi<strong>pe</strong>line, the capital cost of dewatering<br />

the <strong>slurry</strong> to recover the minerals in a solid state combine to define the design criteria of<br />

the pi<strong>pe</strong>line and the physics of the <strong>slurry</strong>.<br />

Corrosion, erosion, and abrasion are very expansive problems that must be taken into<br />

account when designing <strong>slurry</strong> pi<strong>pe</strong>lines. Water hammer can be very destructive and appropriate<br />

protection is recommended during the design as well o<strong>pe</strong>ration phases.<br />

Instrumentation of pi<strong>pe</strong>lines by the use of modern SCADA <strong>systems</strong> has become very<br />

important in the last 20 years to control stations, measure s<strong>pe</strong>ed of flow of <strong>slurry</strong>, monitor<br />

leakage, and detect problems of sedimentation, surges, and transients.<br />

There was great interest in the 1980s following the energy crisis of the 1970s to<br />

convert thermal plants for the direct combustion of coal–water and coal–oil <strong>slurry</strong> mixtures,<br />

particularly when economists were forecasting prices of $50 <strong>pe</strong>r barrel. Unfortunately,<br />

this interest in burning <strong>slurry</strong> mixtures dwindled in the 1990s. By the year 2000,<br />

some plants had started to burn tar–water emulsions such as the Venezuelan Orimulsion<br />

and some old coal thermal plants were converted to burn to this interesting mixture.<br />

There remain many concerns about transporting <strong>slurry</strong> mixtures in ships, as the solids<br />

would not float and their recovery would be almost impossible in the case of an accident.<br />

11-1 BAUXITE PUMPING<br />

Bauxite is a prime source of aluminum. Pumping of bauxite slurries was discussed in Section<br />

5-11-1, particularly the work of Want et al. (1982), with a worked example.<br />

11.1


11.2 CHAPTER ELEVEN<br />

11-2 GOLD TAILINGS<br />

Gold tailings at high concentration flow as non-Newtonian mixtures. The work of Sauermann<br />

(1982) in this field was examined in Section 5-12. Gold tailings pi<strong>pe</strong>lines are of a<br />

small diameter, because of the limited quantities of materials to pump. In 2001, a new<br />

tailings pi<strong>pe</strong>line was commissioned at Homestake Eskay Creek, in the north of British<br />

Columbia, Canada. The pi<strong>pe</strong>line is 6500 m long (4 miles). The inner diameter of the<br />

pi<strong>pe</strong>line was sized at 2.5� and the flow was in the range of 20 m 3 /hr (88 US gpm). The<br />

flow was of a homogeneous Bingham plastic rheology. A Wirth positive displacement<br />

pump was installed and sized at 14 MPa (2000 psi).<br />

11-3 COAL SLURRIES<br />

Coal is an important fuel for power plants. Its transportation in the form of <strong>slurry</strong> has received<br />

considerable attention since the successful construction of the Black Mesa Pi<strong>pe</strong>line<br />

(Figure 11-1). In fact, one of the longest <strong>slurry</strong> pi<strong>pe</strong>lines is the ETSI coal pi<strong>pe</strong>line, built in<br />

1979. It spans a distance of 1670 km (1036 miles), uses a 965 mm (38 in) pi<strong>pe</strong>, and transports<br />

23 million metric tons/year (25 US tons/year). In Russia, the Siberian coal pi<strong>pe</strong>line is<br />

260 km (163 mi) long and transports 4 million tons of coal a year from Siberian mines.<br />

11-3-1 Size of Coal Particles<br />

There has been considerable research on coal transportation in the form of <strong>slurry</strong>. The Euro<strong>pe</strong>an<br />

researchers favored transporting coarse <strong>slurry</strong> over relatively short distances. The<br />

FIGURE 11-1 The Black Mesa pi<strong>pe</strong>line was one of the first long-distance coal <strong>slurry</strong><br />

pi<strong>pe</strong>lines.


SLURRY PIPELINES<br />

U.S. engineers favored fine and well-ground coal over much longer distances. As a result,<br />

all kinds of schemes ranging from coal as coarse as 50 mm (2 in) to very fine with a diameter<br />

smaller than 1 mm (0.039 in) have been studied.<br />

Brookes and Dodwell (1985) reported that coal up to a diameter of 150mm (6 in) at a<br />

weight concentration of 35% was pum<strong>pe</strong>d to the Hammersmith Power Station. Coal with a<br />

top size of 50 mm (2 in) was pum<strong>pe</strong>d at a weight concentration of 60% and a s<strong>pe</strong>ed of 2.5<br />

m/s (8.2 ft/sec). This technique using water as a carrier was widely used for short haulage,<br />

such as at the Loveridge mine in the United States and the Hansa mine in Germany, for distances<br />

up to 6.5 km (4 miles) and for vertical lifts of 600 m (2000 ft). This reduces the dewatering<br />

cost, but is achieved at a high power consumption of 0.8 to 1.0 kWh/tonne-km.<br />

Hydromechanically extracted coal (hydrocal) may be transported in a natural state,<br />

without grinding, over short distances. Particle sizes may range from 0–60 mm (0–2.4 in).<br />

The density of coal varies de<strong>pe</strong>nding on the moisture content. An average s<strong>pe</strong>cific gravity<br />

of 1.35 is often used in calculations.<br />

Leninger et al. (1978) classified the ty<strong>pe</strong>s of coal transported as <strong>slurry</strong>:<br />

� Power plant coal<br />

� Coking coal<br />

� Flotation tailings<br />

� Hydrocoal or hydromechanically extracted coal<br />

In long-distance pumping of coal, Lenninger et al. (1978) indicated that the feeding<br />

plant and discharge dewatering plant might exceed 30% of the cost of the pi<strong>pe</strong>line. This is<br />

due to the equipment associated with screening, crushing, thickening, filtering, and mechanical<br />

dewatering.<br />

Power plant coal is usually ground and dried before being fed to a boiler for burning.<br />

Moisture content must be reduced to less then 10% for pro<strong>pe</strong>r handling and burning of<br />

coal. This is not an easy task, as coal has a natural tendency to retain water. To avoid pollution<br />

problems, the fly ash content must be low.<br />

Klose and Mahler conducted tests on the critical s<strong>pe</strong>ed of coal with a top size of 10<br />

mm (� 0.4 in) as a function of weight concentration and pi<strong>pe</strong> diameter. Results are presented<br />

in Figure 11-2.<br />

11-3-2 Degradation of Coal During Hydraulic Transport<br />

11.3<br />

Hydrotransport of coking coal must be done in such a way as to avoid excessive degradation<br />

and oxidation. Lenninger et al. (1978) indicated that the finer the particles of coking<br />

coal, the worse the deterioration by oxidation. They suggest that coking coal be transported<br />

with the top size not exceeding 3.15 mm (0.124 in).<br />

Flotation slimes from coal circuits typically have a top size of 0.75 mm (0.03 in).<br />

One feature of coarse coal is that it tends to break up into smaller particles as it is<br />

pum<strong>pe</strong>d over long distances. The degradation of coal during hydraulic transportation<br />

must be taken in account when calculating the deposition velocity as well as the pressure<br />

drop. Friable and coarse lignite degrades rapidly. Particle degradation is accentuated<br />

when there are numerous bends and valves in the pi<strong>pe</strong>line. Shook et al. (1979) reported on<br />

lab tests on coal degradation. They concluded that:<br />

� Particle degradation in recirculating loops is similar to rod and ball milling processes.<br />

� There is a time factor to consider. There may be a substantial initial degradation with<br />

friable coals that decreases exponentially with time.


11.4 CHAPTER ELEVEN<br />

Critical S<strong>pe</strong>ed (m/s)<br />

2.0<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

1.0<br />

0.3 0.4 0.5<br />

Weight Concentration<br />

� Lignite coal degrades at higher im<strong>pe</strong>ller diameter tip s<strong>pe</strong>eds.<br />

� For bitumous coal, degradation is linear with relatively small particles, but degradation<br />

is lower with coarser particles.<br />

� Therefore, degradation is not uniform from one grade to another. Degradation can result<br />

in a bimodal distribution of coarse and fine particles<br />

The degradation of coal in a pi<strong>pe</strong>line is an interesting phenomenon. Its physics was<br />

used to understand the degradation of oil sands. There are other slurries in nature, such as<br />

the Canadian oil sands, that are sometimes pum<strong>pe</strong>d in sizes as large as 50 mm (2 in).<br />

Mixed with a s<strong>pe</strong>cial solvent and pum<strong>pe</strong>d over long distances, the tar–sand balls break up<br />

during transportation and the tar separation is facilitated. This ingenious approach, develo<strong>pe</strong>d<br />

in Canada, has allowed a shift in technology from hot steam treatment to warm water<br />

and much less ex<strong>pe</strong>nsive processes. It is not farfetched to say that the work on coal<br />

degradation started by Pi<strong>pe</strong>lin et al. (1966) on coal as early as the 1960s and continued by<br />

a number of scientists such as Shook et al. (1979) was the basis of a new technology for<br />

oil–sand separation.<br />

Pi<strong>pe</strong>lin et al. (1966) reported that the degradation of coal in pumps was a function of<br />

the im<strong>pe</strong>ller tip s<strong>pe</strong>ed or tangential s<strong>pe</strong>ed at the tip diameter raised to the power of 2.5.<br />

Corrosion control is important in the case of coal <strong>slurry</strong> pi<strong>pe</strong>lines because of the presence<br />

of sulfur in coal.<br />

Ercolani et al. (1988) conducted extensive closed-loop tests on fine coal and water<br />

mixtures at shear rates of 20/sec, 50/sec, and 80/sec. They concluded that pumping at<br />

power rates in the range of 0.1 to 2.0 W/kg (0.045–0.90 W/lb) did not seem to affect the<br />

degradation of coal under mechanical stress.<br />

11-3-3 Coal–Magnetite Mixtures<br />

NB400 (16")<br />

NB300 (12")<br />

NB200 (8")<br />

FIGURE 11-2 Critical s<strong>pe</strong>ed of coal with top size of 10 mm (0.394 in) as a function of<br />

weight concentration and pi<strong>pe</strong> diameter. (After Klose and Mahler, 1982.)<br />

Buckwalter (1977) investigated the transportation of very coarse coal (>50 mm or >2 in)<br />

in a magnetite-based water mixture. The magnetite consists of very fine particles but they<br />

7<br />

6<br />

5<br />

4<br />

3<br />

Critical s<strong>pe</strong>ed (ft/sec)


are heavier than coal. When magnetite is mixed with water, this mixture becomes the effective<br />

carrier liquid in which the >50 mm (>2 in) coal particles can float.<br />

In Chapter 4, it was clearly explained that the difference in density (or s<strong>pe</strong>cific gravity)<br />

between the carried solids and the carrier liquid was an important parameter in friction<br />

loss calculations. This is, in basic terms, the concept of using a heavy medium (water and<br />

magnetite) as a carrier for coarse solids.<br />

A circuit that uses magnetite must have a recovery system at the end of the pi<strong>pe</strong>line.<br />

Since magnetite has very strong ferromagnetic pro<strong>pe</strong>rties, it is first screened from coarse<br />

coal, and then separated from crushed coal (that resulted from deterioration during pumping),<br />

using magnetic separators. The recovered magnetite is then mixed with water at a<br />

high volumetric concentration and pum<strong>pe</strong>d via a dedicated pi<strong>pe</strong>line and a positive displacement<br />

back to the starting feed station of the <strong>slurry</strong> pi<strong>pe</strong>line. It is then stored in s<strong>pe</strong>cial<br />

storage tanks with agitator mixers.<br />

To avoid the use of many booster stations in long pi<strong>pe</strong>lines. The lockhop<strong>pe</strong>r may be<br />

used for coarse coal (>50mm or >2 in; see Figure 9-17).<br />

11-3-4 Chemical Additions to Coal–Water Mixtures<br />

S<strong>pe</strong>cial chemicals such as xanthan gum in levels of 200–600 ppm can be used as a stabilizer<br />

to prevent settling of coal slurries and to prevent the formation of hard-packed beds<br />

during hydrotransport. Miller and Hoyt (1988) recommended Pfizer Flocon 4800C as a<br />

very economical additive to coal–water mixtures.<br />

Morway (1965) obtained a patent for using a hydrocarbon oil with a small <strong>pe</strong>rcentage<br />

of an imidazoline surfactant to coat coal particles uniformly. After adding this mixture,<br />

the coated coal can be mixed with water. The water weight concentration can be reduced<br />

to 20%. This <strong>slurry</strong> with low overall moisture is easier to heat at the final discharge point<br />

prior to combustion than plain coal–water mixtures.<br />

Bomberger (1965) proposed hexametaphosphate and sulfite as corrosion inhibitors for<br />

coal slurries in steel pi<strong>pe</strong>lines.<br />

11-3-5 Coal–Oil Mixtures<br />

SLURRY PIPELINES<br />

11.5<br />

The vast majority of slurries consist of water and solids; however, variations on this are<br />

being implemented, es<strong>pe</strong>cially in the transportation of coal. In the case of thermal plants,<br />

slurries of water and coal are difficult to burn, and a complete process of dewatering is<br />

needed to separate the coal from water. To rectify this situation, proposals have been<br />

made to use heavy crude oils instead of water to transport and burn coal.<br />

The viscosity of a heavy oil combined with the degradation of coal during pi<strong>pe</strong>line<br />

transportation ultimately leads to non-Newtonian flows. The rheology of such mixtures de<strong>pe</strong>nds<br />

on particle size, tem<strong>pe</strong>rature, concentration, and the quality of the coal and the carrier<br />

oil. With the worsening political situation that started in 2001, coal–oil mixtures may see<br />

more and more applications. Kreusing and Franke (1979) recommended the use of coal<br />

particles smaller than 5 mm (0.2 in) as a fuel for the blast furnace. To maintain the viscosity<br />

of a coal-oil mixture in a range that is pumpable, the <strong>slurry</strong> may have to be warmed to<br />

50° C (122° F) [fig 11-3]. By using heat, Kreusing and Franke (1979) managed to maintain<br />

a <strong>slurry</strong> viscosity in the range of 13–40 Pa·s (whereas the viscosity of water is 0.001 Pa·s).<br />

A viscosity of 40 Pa·s is an up<strong>pe</strong>r limit allowable for use in centrifugal pumps.<br />

In the case of carbonized lignite, Kreusing and Franke (1979) achieved a maximum<br />

weight concentration of 34%, and with brown coal they achieved a maximum weight concentration<br />

of 41–46%, de<strong>pe</strong>nding on the solid particle size. Brown coal is often a low-


11.6 CHAPTER ELEVEN<br />

calorific coal and is difficult to export. Kreusing and Franke achieved a maximum weight<br />

concentration of 60% with mineral coal. They concluded that it was possible to use coal<br />

for a maximum energy substitution of oil of 52%. Since coal is chea<strong>pe</strong>r than oil, this is not<br />

a negligible result (Figures 11-3 and 11-4).<br />

11-3-6<br />

Dewatering Coal Slurry<br />

At the discharge point of a coal <strong>slurry</strong> pi<strong>pe</strong>line, water must be removed because coal cannot<br />

be burned at such high water contents. It is essential to establish certain criteria for the<br />

design of a dewatering station:<br />

� Size distribution<br />

� Water content<br />

� End use<br />

� Rheology<br />

� Volumetric concentration<br />

� Suitability of recovered water for further use<br />

� Storage, stockpiling, or ship loading<br />

� Coal degradation during transport<br />

FIGURE 11-3 Viscosity of coal–oil <strong>slurry</strong> mixture. (From Kreusing and Franke, 1979.<br />

Reprinted with <strong>pe</strong>rmission of BHR Group.)


SLURRY PIPELINES<br />

11.7<br />

FIGURE 11-4 Viscosity as a function of tem<strong>pe</strong>rature for coal–oil mixture. (From Kreusing<br />

and Franke, 1979. Reprinted with <strong>pe</strong>rmission of BHR Group.)<br />

Dewatering may be done mechanically via filter presses, centrifuges, and screens. Because<br />

the efficiency of dewatering often de<strong>pe</strong>nds on the particle size distribution, screening<br />

is strongly recommended prior to dewatering.<br />

Leninger et al. (1978) reported that in the case of coal with a top size of 10 mm (0.4<br />

in), it was very difficult to use mechanical dewatering devices to reduce moisture below<br />

10.6 %, even using steam with vacuum filtering.<br />

For coals with a top size of 3.15 mm (0.125 in), Leninger et al. (1978) suggested using<br />

two-stage cyclones with the second stage connected to the overflow of the first. The underflow<br />

from the first stage as well as from the second stage should be fed into solid-bowl<br />

centrifuges to reduce the residual moisture content to 17.3%. The use of hydrocyclones<br />

was also discussed by Abbot (1965).<br />

Coal with a top size of 2.4 mm (0.09 in) can be dewatered using two-stage cyclones as<br />

well as solid-bowl centrifuges, but the overall moisture is only reduced to 19%.<br />

For ultra-fine coal with a top size of 1.4 mm (0.055 in), as in the case of the Ohio and<br />

the Black Mesa pi<strong>pe</strong>lines, mechanical dewatering must be followed by thermal drying.<br />

One way to reduce the cost is to use the waste gas of the thermal plant to dry the coal.


11.8 CHAPTER ELEVEN<br />

11-3-7 Ship Loading Coarse Coal<br />

The endpoint of a coal <strong>slurry</strong> pi<strong>pe</strong>line may be a thermal power plant or a ship loading facility<br />

(Figure 11-5) for export. Faddick (1982) reviewed some concepts for loading coarse<br />

coal onto ships:<br />

� Submarine pi<strong>pe</strong>line<br />

� Vertical riser<br />

� Single-point mooring system<br />

� Flexible hose system<br />

The single point mooring (SPM) system is the most economical and feasible concept.<br />

In 1971, the Marcona Corporation installed at Walpi<strong>pe</strong>, New Zealand the first successful<br />

SPM for ship loading of iron sand <strong>slurry</strong>. Reaching a large vessel with <strong>slurry</strong> is not always<br />

easy. Many ships require very deep ports or must stay out in deep waters to be<br />

loaded from a mooring point, as is the case with oil tankers. In United States ports, large<br />

carriers with drafts greater than 18–20 m (65–70 ft) have to be reached at great distances<br />

due to the relative shallowness of most American ports.<br />

Submarine pi<strong>pe</strong>lines are widely used in the oil and gas industry. For slurries, difficult<br />

accesses with possible sedimentation of coarse particles tend to be discouraging. Submarine<br />

<strong>slurry</strong> pi<strong>pe</strong>lines are typically limited to nonsettling slurries for tailings disposal, or to<br />

relatively short distances.<br />

High-density polyethylene (HDPE) pi<strong>pe</strong>s are lighter than water and can be floated.<br />

There are no records of using HDPE pi<strong>pe</strong>s with coarse coal and it is unknown whether<br />

they can be used as submarine pi<strong>pe</strong>s with this <strong>slurry</strong>. Flexible hoses are very ex<strong>pe</strong>nsive in<br />

large diameter sizes in excess of 250 mm (10 in) NB (normal bore). Adams (1986) did indicate<br />

that the use of polyethylene pi<strong>pe</strong> is limited to solids with a maximum diameter of<br />

9.5 mm [3/8 in], which would certainly not make these pi<strong>pe</strong>s suitable for fairly coarse<br />

solids.<br />

In Chapter 4, the Russian’s work on coarse coal friction losses and deposition velocity<br />

was presented in Section 4-4-3. The Russian equations are useful as an alternative to complex<br />

stratified models.<br />

The density of coal varies de<strong>pe</strong>nding on the moisture content. An average s<strong>pe</strong>cific<br />

gravity of 1.35 is often used in calculations.<br />

11-3-8 Combustion of Coal–Water Mixtures (CWM)<br />

In the 1980s, considerable research was conducted in the United States on converting<br />

diesel engines to burning coal or its derivatives. The interest in coal fired diesel engines<br />

died away in the 1990s when the price of oil dwindled to $12 a barrel. With the threat<br />

of a new energy crisis at the turn of the 21st century, interest in coal fired diesel engines<br />

may revive. The most promising schemes required gasification of coal into a combustible<br />

gas. In an effort to bypass the gasification, plant schemes were proposed to<br />

burn coal as <strong>slurry</strong> in a diesel engine. Tests showed that coal would wear out piston<br />

rings, engine liners, etc. The concept required s<strong>pe</strong>cial construction materials, as is often<br />

the case with slurries.<br />

In an effort to bypass the problems of using pulverized coal in a <strong>slurry</strong> form with solid<br />

pistons, researchers have investigated liquid piston engines. The idea of a liquid piston<br />

engine was pioneered in the 19th century in the United Kingdom by Humphrey Engines.<br />

In an effort to bypass the problems of wear due to exploding a <strong>slurry</strong> mixture against sol-


Product Distribution Unit<br />

Slurry Marine Hoses<br />

FIGURE 11-5 Ship loading of coal <strong>slurry</strong>. (From Faddick, 1982. Reproduced with <strong>pe</strong>rmission<br />

of BHR Group.)<br />

11.9<br />

Seal Water Wash System<br />

Slurry Pi<strong>pe</strong>line<br />

End Manifold<br />

Slurry Loading Arm


11.10 CHAPTER ELEVEN<br />

id pistons, Abulnaga (1990) develo<strong>pe</strong>d a liquid piston engine and obtained an Australian<br />

patent.<br />

The Abulnaga engine features a s<strong>pe</strong>cial Darrieus or Savonius rotor between two cylinders.<br />

Essentially, this means that the liquid pistons are formed as a column of liquid that<br />

oscillates between two cylinders. The great advantage of the Darrieus or Savonius rotors<br />

is that they maintain the same direction of rotation irres<strong>pe</strong>ctive of the oscillation direction<br />

of the column.<br />

Typical coal–water mixtures (CWMs) for direct combustion consist of a weight concentration<br />

of 70% coal and 30% water. Prior to combustion, it is important to degrade the<br />

coal <strong>slurry</strong> mixture by applying hot air to accelerate the evaporation of water (Garbett and<br />

Yiu, 1988).<br />

The concept of direct combustion of CWM in gas turbines has been proposed in the<br />

literature. There remain many unknowns, particularly as to the erosion of the blades from<br />

fly ash.<br />

11-3-9 Pumping Coal Slurry Mixtures<br />

Hughes (1986) described the development of positive displacement pumps for the Siberian<br />

coal pi<strong>pe</strong>line Belovo-Novosibirsk. This pi<strong>pe</strong>line is 256 km (160 mi) long. It transports<br />

heavily concentrated water–coal <strong>slurry</strong>. This pi<strong>pe</strong>line features one main pump station and<br />

two booster stations. Each pump station features single-acting triplex Ingersoll Dresser<br />

pumps. S<strong>pe</strong>cial 100 bar (1,470 psi) gate valves were manufactured in sizes of 200 mm (8<br />

in) and 350 mm (14 in).<br />

Vanderpan (1982) recommended the use of Ni-hard as a material to cast the im<strong>pe</strong>llers<br />

and liners of coal handling <strong>slurry</strong> pumps. For certain high pH applications due to acidic<br />

water, or in the case of high-salt mixtures, s<strong>pe</strong>cial high-alloy irons may be used instead of<br />

Ni-hard.<br />

The use of centrifugal pumps in series is usually limited to a discharge pressure of 4.2<br />

MPa (or 600 psi). This may be suitable for puming coarse coal up to a distance of 50 km<br />

(30 mi).<br />

11-4 LIMESTONE PIPELINES<br />

Limestone is an important material. It was used thousands of years ago to build the colossal<br />

pyramids of Egypt and is used today to manufacture cement and concrete. Many derivatives<br />

of limestone are used as fertilizer, for alkalination of chemical processes, and as<br />

a pollution control substance used to absorb sulfur dioxide pollutants in flue gas desulphurization.<br />

In Chapter 1, a number of limestone pi<strong>pe</strong>lines are listed. The pi<strong>pe</strong>line focused on in<br />

this segment is the Gladstone pi<strong>pe</strong>line of the Queensland Cement and Lime Company in<br />

Australia, which started o<strong>pe</strong>ration in 1979. Venton (1982) described the pi<strong>pe</strong>line in great<br />

detail and diagramed examples of a practical design for a cement plant.<br />

The Gladstone pi<strong>pe</strong>line is 24 km long and is located 400 km north of Brisbane. The reserves<br />

of limestone are overlaid by deposits of clay suitable for a clinker cement plant. A<br />

ship loading facility was built in the Gladstone harbor in order to transport the lime to<br />

Brisbane (Figure 11-6).<br />

At the discharge point of the limestone pi<strong>pe</strong>line, the <strong>slurry</strong> is dewatered by a thermal<br />

drying processes. The lime is then transported in a powder form. To reduce the relatively


SLURRY PIPELINES<br />

11.11<br />

FIGURE 11-6 The Gladstone limestone pi<strong>pe</strong>line in Queensland, Australia. (From Venton,<br />

1982. Reprinted with <strong>pe</strong>rmission of BHR Group.)


11.12 CHAPTER ELEVEN<br />

high cost of thermal drying, mechanical dewatering with filter presses is used. The lime is<br />

therefore delivered in a semiwet state, but this is acceptable for a cement plant. The filter<br />

presses reduce the water content from 36% to 17% moisture. In the actual cement plant,<br />

waste heat from the kiln off-gases is then used for further drying.<br />

In a limestone pi<strong>pe</strong>line project, the <strong>slurry</strong> plant is located at the quarry and needs to be<br />

fitted with a milling or grinding circuit. Water also must be available on-site to be blended<br />

with the ground limestone. Some of this water may be available locally; however, if<br />

the pi<strong>pe</strong>line is relatively short, it may be possible to return water from the pi<strong>pe</strong>line discharge<br />

point.<br />

Processes of <strong>slurry</strong> preparation are described in Chapter 7. Limestone pi<strong>pe</strong>lines typically<br />

o<strong>pe</strong>rate in a range of 50–60% weight concentration. If other ingredients such as clay<br />

are in the <strong>slurry</strong>, a small pump test loop is recommended on-site to monitor the composition<br />

and concentration.<br />

In the case of the Gladstone pi<strong>pe</strong>line, the viscosity was in the range of 10 mPa·s (1 cP)<br />

at a weight concentration of 56%, 20 mPa·s (2 cP) at a weight concentration of 60%, but<br />

rose sharply toward a viscosity of 70 mPa·s (7 cP) at a weight concentration of 68%. The<br />

yield stress was in a range of 8–14 Pa (Figure 11-7). The laminar to turbulent velocity in a<br />

200 mm (8 in) pi<strong>pe</strong>line was predicted to be in the range of 1–1.3 m/s at a weight concentration<br />

of 62–64%.<br />

The Gladstone pi<strong>pe</strong>line uses Wilson-Snyder positive displacement pumps. At the<br />

weight concentration of 60–64%, the <strong>slurry</strong> acted as a Bingham mixture, with non-Newtonian<br />

viscosity characteristics. However, it did feature clay, sand, and iron, as the materials<br />

were formulated for the manufacture of clinker cement. The pi<strong>pe</strong>line o<strong>pe</strong>ration s<strong>pe</strong>ed<br />

was maintained at 2 m/s and the pressure drop was around 300 kPa/km. API 5LX steel<br />

with a high yield strength was used. Corrosion rates as high as 0.25 mm/year were measured<br />

during the initial phase of o<strong>pe</strong>ration of the pi<strong>pe</strong>line.<br />

Pertuit (1985) reported that during the first two years of o<strong>pe</strong>ration problems of o<strong>pe</strong>ration<br />

included:<br />

� Severe knocking and vibration of mainline pumps<br />

� Short life of gland packing and piston scouring of the positive displacement pumps<br />

These problems were eventually solved.<br />

The extremely high rate of corrosion was unex<strong>pe</strong>cted, since the lab reports suggested<br />

a design for a low corrosion rate of 0.076 mm/year. Venton (1982) reported that the<br />

o<strong>pe</strong>rators brought down the rate of corrosion by adding inhibitors to the <strong>slurry</strong> composition.<br />

11-5 IRON ORE SLURRY PIPELINES<br />

Iron ore is critical to our modern economy. A number of iron ore <strong>slurry</strong> pi<strong>pe</strong>lines have<br />

been constructed since the 1960s (see Table 1-9 for s<strong>pe</strong>cific examples). One of the most<br />

famous is the SAMARCO pi<strong>pe</strong>line in Brazil, which is 390 km (245 mi) long.<br />

In order to understand its rheology, Thomas (1976) conducted tests on iron ore at a<br />

volumetric concentration of 24% and with a solids diameter d50 = 40 �m. The tests were<br />

conducted in 150 mm (6 in) and 200 mm (8 in) pi<strong>pe</strong>. The head loss in meters of water <strong>pe</strong>r<br />

meter of pi<strong>pe</strong> length was derived as<br />

� = KV x y DI (11-1)


SLURRY PIPELINES<br />

11.13<br />

FIGURE 11-7 Rheology of the Glasdtone limestone <strong>slurry</strong>. (From Venton, 1982. Reprinted<br />

with <strong>pe</strong>rmission of BHR Group.)


11.14 CHAPTER ELEVEN<br />

where<br />

K = 5228<br />

x = 1.77<br />

y = –1.18<br />

The equation of Thomas does not agree well with ex<strong>pe</strong>rimental data published by<br />

Hayashi et al. (1980). However, this may be due to the difference in particle size distribution.<br />

Lokon et al. (1982) conducted further tests at the Melbourne Institute of Technology<br />

in Australia and derived the following equation for the pressure loss gradient in Pa/m:<br />

i m = KV x D I y Cv z (11-2)<br />

where<br />

im = friction gradient of the mixture<br />

k = 54.9<br />

x = 1.63<br />

y = –1.42<br />

z = 0.35<br />

Lokon et al. indicated that their power law compared well with commercial pi<strong>pe</strong>lines.<br />

Their data was based on iron ores pum<strong>pe</strong>d with solids in a size range of 30–60 �m (mesh<br />

325–250). Obviously, this was the range of nonsettling slurries. Pressure losses are presented<br />

in Figure 11-8.<br />

Example 11-1<br />

Using the Lokon equation, determine the pressure for an iron ore pi<strong>pe</strong>line under the following<br />

conditions:<br />

� Pi<strong>pe</strong>line inner diameter is 175 mm<br />

� Volumetric concentration is 33%<br />

� Flow rate is 48 L/s<br />

� Pi<strong>pe</strong>line length is 50 km<br />

Solution in SI Units<br />

flow area A = 0.25 × 0.175 2 = 0.024 m 2<br />

flow s<strong>pe</strong>ed V = Q/A = 0.048/0.024 = 2 m/s<br />

im = KV x y z DI Cv<br />

i m = 54.9 × 2 1.63 × 0.175 –1.42 × 0.33 0.35<br />

im = 54.9 × 3.095 × 11.88 × 0.6784<br />

im = 1369 Pa/m<br />

Klose and Mahler (1982) measured the critical s<strong>pe</strong>ed of iron ore <strong>slurry</strong> with particles<br />

size in the range of 1 to 2 mm (0.04–0.08 in). However, due to the high density of iron oxide<br />

(SG = 5.0) critical s<strong>pe</strong>ed as high as 3.5 m/s (11.5 ft/sec) were recorded (Figure 11-9).<br />

To design economic pi<strong>pe</strong>lines, Klose and Mahler suggested the addition of s<strong>pe</strong>cial chemical<br />

additives that can reduce the critical s<strong>pe</strong>ed of the mixture, despite a slight rise in the<br />

pressure drop at these lower s<strong>pe</strong>eds (Figure 11-10).


5000<br />

4000<br />

3000<br />

2000<br />

1000<br />

300<br />

600<br />

500<br />

SLURRY PIPELINES<br />

PRESSURE GRADIENT (kpa/Km)<br />

� IRON ORE C V = 24.9<br />

� IRON ORE C V = 26.6<br />

� IRON ORE C V = 28.1<br />

� IRON ORE C V = 30.2<br />

� IRON ORE C V = 31.3<br />

� WATER<br />

VELOCITY (m/s)<br />

0.5 1 2 3 4 5<br />

11.15<br />

FIGURE 11-8 Pressure losses for iron ore oxides in the range of 30–60 �m (mesh 325-250).<br />

(From Lokon et al., 1982. Reprinted with <strong>pe</strong>rmission of BHRA Group.)<br />

Taconite is a very important source of iron in the United States. It is a form of iron<br />

sand found in the Mesabi range of Minnesota, as well as in Manitoba and Ontario, Canada.<br />

The Shilling Mining Review (1981), in an editorial article, reported on the pumping of<br />

taconite tailings using 20 in × 18 in (500 mm × 450 mm) Warman tailings pumps sized to<br />

a pressure of 350 psi. The pumps were installed in six stages.<br />

Taconite tailings are considered coarse sand and must be pum<strong>pe</strong>d in a range of s<strong>pe</strong>eds<br />

of 3.4–4.3 m/s (11–14 ft/s). Rubber-lined pi<strong>pe</strong>s are used. HDPE pi<strong>pe</strong>s are subject to very<br />

fast wear and are not used for tailings disposal. Taconite tailings are typically pum<strong>pe</strong>d at a<br />

weight concentration of 35%. The use of s<strong>pe</strong>cial flocculants in modern, efficient thickeners<br />

allows pumping up to a weight concentration of 45%.<br />

The SAMARCO pi<strong>pe</strong>line in Brazil is one of the longest ever built to transport iron ore<br />

oxides and features 500 mm (20 in) and 457 mm (18 in) pi<strong>pe</strong> sections over a distance of<br />

400 km (250 miles). Start-up occurred in 1977 and it is ex<strong>pe</strong>cted to remain in o<strong>pe</strong>ration<br />

for 40 years (Weston, 1985).<br />

Another long pi<strong>pe</strong>line to transport iron ore oxide is the La Perla-Hercules pi<strong>pe</strong>line in<br />

Mexico, with an overall length of 382 km (239 miles). The pi<strong>pe</strong>line features one main and<br />

one booster pump station with single-acting triplex plunger pumps (Thompson, 1995).


11.16 CHAPTER ELEVEN<br />

Velocity, v (m/s)<br />

Concentration, c v<br />

FIGURE 11-9 Critical s<strong>pe</strong>ed of iron ore oxides with particle size in the range of 1 to 2 mm<br />

(0.04–0.08 in). (After Klose and Mahler, 1982. Reprinted with <strong>pe</strong>rmission of BHR Group.)<br />

11-6 PHOSPHATE AND PHOSPHORIC ACID<br />

SLURRIES<br />

Phosphate is a very important source of fertilizer for agriculture and is mined in large<br />

quantities in the United States, Morroco, Egypt, South Africa, China, and other countries.<br />

Phosphate rock is sometimes transported in a pre-milled state over a relatively<br />

short distance—a few kilometers or miles. In Florida, a method of transporting phosphate<br />

rock while mining it in a very similar fashion to dredging a river using a pump<br />

has been develo<strong>pe</strong>d. Large phosphate matrix pumps driven by diesel engines are available<br />

on baseplates. These are relocated from site to site as the phosphate matrix field is<br />

mined out.<br />

Tillotson (1953) described the phosphate matrix in Polk and Hillsborough counties.<br />

About 5120 km 2 (2000 mi 2 ) contained high grades of phosphate. Eight million short tons<br />

of saleable phosphate <strong>pe</strong>bbles were produced annually. Tillotson described the phosphate<br />

matrix as an unconsolidated mixture of clay smaller than 1 mm (0.04 in), silica sand, and<br />

phosphate rock of a much larger size. This mixture ranged in size from rocks as large as<br />

150 mm (6 in) to as small as 38 �m (400 mesh). Because phosphate matrix pumps have to<br />

handle lumps as large as 150 mm (6 in), they tend to be large with 500 mm (20 in) suction<br />

and discharge sizes.


Pressure loss Dp (10 5 Pa/1000 ml)<br />

22<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

vv c<br />

DN 200<br />

c = v = 0.64<br />

c v<br />

v vc<br />

c<br />

Once the phosphate matrix is pum<strong>pe</strong>d or transported, it is processed in a s<strong>pe</strong>cial phosphate<br />

rock treatment plant. Nordin (1982), of the Phospnate Development Corporation<br />

Ltd., described how each year a South African plant produces approximately three million<br />

tons of phosphate rock from foskorite and pyroxenite ores. The ore was then classified<br />

through flotation, thickening, and filters before being stockpiled. The result was fine<br />

gray-white crystalline powder of mineral apatite, with a 36.5% P 2O 5 content, a solid s<strong>pe</strong>cific<br />

gravity of 3.17, and d 50 � 106 �m. The hardness of the apatite was measured at 5.0<br />

on the Mohr scale.<br />

Nordin (1982) reported that milled phosphate rock is easy to pump in a weight concentration<br />

of 30–70%. He conducted tests on a 100 mm (4 in) loop and obtained the values<br />

of critical velocity shown in Table 11-1, where d 70 � 75 �m. He recommended<br />

pumping at 0.3 m/s (1 ft/s) above critical s<strong>pe</strong>ed (Table 11-2).<br />

11-6-1 Rheology<br />

4%<br />

4%<br />

3% 3%<br />

v vcc<br />

SLURRY PIPELINES<br />

2% 2%<br />

0<br />

0 1.0 2.0 3.0 4.0 5.0 6.0<br />

Velocity, v (m/s)<br />

11.17<br />

FIGURE 11-10 Reduction of critical s<strong>pe</strong>ed of iron ore oxides with particle size in the range<br />

of 1 to 2 mm (0.04–0.08 in) by the use of s<strong>pe</strong>cial chemical additives. (After Klose and Mahler,<br />

1982. Reprinted with <strong>pe</strong>rmission of BHR Group.)<br />

Landel et al. (1963) investigated the rheology of a bimodal (fine and coarse) distribution<br />

of phosphate ore. They reported that in certain cases the finer particles act as a carrier for


11.18 CHAPTER ELEVEN<br />

TABLE 11-1 Critical and Recommended S<strong>pe</strong>ed of Pumping Phosphate Rock with d 70<br />

� 75 �m (data from Nordin, 1982)<br />

the coarser solids and that for all intents and purposes the <strong>slurry</strong> may be considered non-<br />

Newtonian. They proposed the following equation for the consistency factor:<br />

Cv �<br />

Cmax<br />

K = � L� 1 – � –2.5<br />

where<br />

K = fluid consistency index (Pa·s)<br />

C max = maximum solids volumetric concentration<br />

C v = volumetric concentration of solids<br />

Peterson and Mackie (1996) proposed the following equation for phosphate ore:<br />

(11-3)<br />

�0 = B���� (11-4)<br />

Cmax – Cv where B = 13.3.<br />

The data presented by Peterson and Mackie (1996) on the critical s<strong>pe</strong>ed is consistent with<br />

the data from Nordin (1982).<br />

Anand et al. (1986) indicated that the corrosion rate due to Maton phosphate is of the<br />

order of 0.3 mm/year in steel pi<strong>pe</strong>s. A total corrosion and wear allowance of 0.4 mm/year<br />

is suggested by Peterson and Mackie (1996). It is, however, recommended to assume<br />

more wear in the initial dozens of kilometers (miles) in a long pi<strong>pe</strong>line, as particle attrition<br />

and degradation often occur in the initial portion of the pi<strong>pe</strong>line. As the particles become<br />

less sharp, their abrasion of the pi<strong>pe</strong>line decreases (Table 11-3).<br />

11-6-2 Materials Selection for Phosphate<br />

The Miller number of phosphate ore is smaller than 50 (Abulnaga, 2000). This means that<br />

phosphate is suitable for pumping with piston reciprocating pumps.<br />

C v 3<br />

Weight concentration, %<br />

30 40 50 63.5 67 70<br />

Critical velocity m/s 1.30 1.10 0.90 0.85 1.1 1.35<br />

Recommended pumping velocity, m/s 1.60 1.40 1.20 1.15 1.40 1.65<br />

TABLE 11-2 Power Consumption 100 mm (4 in) ID Pi<strong>pe</strong> for Phosphate Rock with<br />

d 70 � 75 �m<br />

Weight concentration, %<br />

30 40 50 63.5 67 70<br />

Power consumption, kW/km 0.196 0.151 0.128 0.112 0.109 0.113<br />

After Nordin (1982).


Pumping phosphoric acid slurries represents a challenge to the manufacturing of <strong>slurry</strong><br />

pumps due to the combination of corrosion and wear in some of the critical circuits<br />

such as flash cooling, filter feed, and gypsum removal. Walker (1993) reported that the<br />

wear life of these pumps can be as low as a few thousand hours.<br />

Traditionally, pumps were lined with rubber or manufactured out of stainless steel.<br />

Rubber linings proved less than optimal. Tearing problems occurred during flash cooler<br />

applications, lowering the life of some components to 3000 hrs. Erratic tearing also decreased<br />

the wear life on filter feeds to as low as 1900 hrs, and local holing (formation of<br />

holes in the liner) decreased wear life to 2300 hrs.<br />

In some res<strong>pe</strong>cts, installing pumps made out of stainless steel is an attractive option<br />

since they can be repaired by welding, but stainless steels are not as hard as abrasionresistant<br />

white iron. A s<strong>pe</strong>cial alloy, which offers as good resistance to corrosion as stainless<br />

steel and hardness as white iron, is Hy<strong>pe</strong>rchrome, develo<strong>pe</strong>d in Australia. Hy<strong>pe</strong>rchrome<br />

was derived from hardfacing weld deposit materials defined in the Australian<br />

Standard as AS2576-1982 Ty<strong>pe</strong> 2. Walker (1993) described an important improvement of<br />

wear life components over comparable stainless steel components.<br />

high chromium content similar to hy<strong>pe</strong>rchrome. The new alloy achieved a service life of<br />

2.5–3 years for an im<strong>pe</strong>ller in gypsum tailings service in a phosphoric acid environment,<br />

whereas the Cd 4MCu material im<strong>pe</strong>ller was badly worn out after 3 months of o<strong>pe</strong>ration.<br />

Both Weir-Warman and KSB-GIW, the largest manufacturers of <strong>slurry</strong> pumps, have<br />

ex<strong>pe</strong>rimented with the use of high chrome alloys with chromes in excess of 30% for <strong>slurry</strong><br />

pumps handling phosphate rock.<br />

The Chevron Pi<strong>pe</strong>line<br />

SLURRY PIPELINES<br />

TABLE 11-3 Example of Phosphate Ore Pro<strong>pe</strong>rties<br />

Pro<strong>pe</strong>rty Fines Coarse Product<br />

Solids s<strong>pe</strong>cific gravity 3.2 3.2 3.2<br />

Freely settled particles packing (%) 40 44 51<br />

Coefficient of sliding friction (�p) 0.53 0.58 0.58<br />

d10 Particle size (�m) 60 50 14<br />

d50 Particle size (�m) 22 92 75<br />

d90 Particle size (�m) 74 150 145<br />

After Paterson (1996).<br />

11.19<br />

Tian et al. (1996) reported the development of a new alloy, a white iron with a very<br />

11-6-3<br />

About 280 km (175 mi) southeast of Salt Lake City, U.S.A. at Vernal, Utah, there is a<br />

large layer of phosphate ore that covers an area of 38,400 km 2 (15,000 mi 2 ) with an estimated<br />

reserve of 700 million short tons (640 million metric tons). Between 1961 and<br />

1986, the ore was transported by truck over a distance of 216 km (135 mi) and then<br />

loaded onto railroad cars. In 1986, a pi<strong>pe</strong>line for the phosphate concentrate was commissioned.<br />

The diameter of the pi<strong>pe</strong>line is 250 mm (10 in). A pi<strong>pe</strong>line feed station with a test<br />

loop and a pump station is located at Vernal, and a booster is located at Richard’s Gap in


11.20 CHAPTER ELEVEN<br />

Wyoming. After covering an overall distance of 150 km (94.3 mi) the pi<strong>pe</strong>line terminates<br />

at Rock Springs, Wyoming. At Vernal, the concentrate is thickened in a s<strong>pe</strong>cial thickener,<br />

then conditioned in three agitator tanks. These large tanks are 15.24 m × 15.24 m (50 ft ×<br />

50 ft), each holding 2000 tons of phosphate. The <strong>slurry</strong> has a weight concentration of<br />

53–60%. Weston and Worthen (19xx) indicated that each agitation tank is fit with a 200<br />

hp mixer. The <strong>slurry</strong> is tested in a 91 m (100 ft) long pump loop prior to being fed to the<br />

pi<strong>pe</strong>line. The pi<strong>pe</strong>line was designed for a flow of 86 L/s (1370 US gpm) and a pressure of<br />

17.915 MPa (2600 psi) <strong>pe</strong>r pump station. The booster station is located at 77 km (48 mi),<br />

halfway along the pi<strong>pe</strong>line.<br />

Two positive displacement Wilson-Snyder pumps were installed at the main pump<br />

station. These were driven by 746 kW (1000 hp) direct current motors. The booster station<br />

is connected to a water pond and draws water on demand to avoid slack flow by providing<br />

additional back pressure in low flow conditions. Choke stations are also provided<br />

for additional back pressure. According to Weston and Worthen, the pi<strong>pe</strong>line used highyield-strength<br />

steel rated at 413 MPa (60,000 psi). An allowance of 2.5 mm (0.1 in) for<br />

corrosion/erosion over a lifetime of 25 years was factored into the design. The thickness<br />

of the pi<strong>pe</strong>line varied between 6.4–12.7 mm (0.25–0.5 in).<br />

The pi<strong>pe</strong>line crossed the Rocky Mountains, so the elevation varied between a low<br />

point of 1676 m (5500 ft) and a high point of 2499 m (8200 ft). To minimize freezing<br />

problems, the pi<strong>pe</strong>line was buried to a depth of 1.8 m (6 ft). There are 12 monitoring or<br />

testing points along the pi<strong>pe</strong>line to monitor for pressure. If freezing or sedimentation develop,<br />

the resultant increase in pressure is automatically detected.<br />

Slurry is pum<strong>pe</strong>d at an average s<strong>pe</strong>ed of 1.5–1.67 m/s (5–5.5 ft/sec) and it takes about<br />

26 hours for the material to be transported from start to finish. The pi<strong>pe</strong>line was designed<br />

to transport 2273 million metric tons (2500 million short tons) of phosphate concentrate<br />

<strong>pe</strong>r year. Initially, it o<strong>pe</strong>rated on a s<strong>pe</strong>cial batch mode with 12 hours of water and 12<br />

hours of <strong>slurry</strong>.<br />

To monitor corrosion, the three following methods are used:<br />

1. Corrosion spools (sacrificial thickness loss)<br />

2. Ultrasonic testing (to measure pi<strong>pe</strong> thickness)<br />

3. Corrosion probes (to measure corrosivity)<br />

Corrosion of phosphate pi<strong>pe</strong>lines is reduced by the use of s<strong>pe</strong>cial inhibitors or by raising<br />

the pH to the alkaline range (alkalination).<br />

11-6-4 The Goiasfertil Phosphate Pi<strong>pe</strong>line<br />

Pertuit (1985) described the Goiasfertil phosphate pi<strong>pe</strong>line. It was constructed in the state<br />

of Goias in Brazil to transport phosphate ore along difficult terrain over a distance of 14.5<br />

km (9 miles) from a mine to an existing railway station in the town of Cataloa. The<br />

pi<strong>pe</strong>line was designed to ship 900,000 metric tons of concentrate <strong>pe</strong>r year over a <strong>pe</strong>riod of<br />

6750 hours. The <strong>slurry</strong> consisted of solids at a weight concentration of 63% to 66%. The<br />

particle distribution consists of 20–25% + 150 microns, and about 25–35% minus 45 microns.<br />

The start-up was in August 1982. The pump station consists of charging centrifugal<br />

pumps, a safety and test loop, and two mainline Wilson-Snyder positive displacement<br />

pumps. They are controlled by variable s<strong>pe</strong>ed drivers and the s<strong>pe</strong>ed is adjusted according<br />

to the flow rate.


11-6-5 The Hindustan Zinc Phosphate Pi<strong>pe</strong>line<br />

Pertuit (1985) described the Hindustan zinc phosphate pi<strong>pe</strong>line, which was commissioned<br />

in late 1983 near the town of Udaipar in India. This pi<strong>pe</strong>line was designed to ship on an<br />

intermittent basis 400 metric tons of phosphate <strong>pe</strong>r year via a 73 mm (2.875 in) diameter<br />

pi<strong>pe</strong> over a distance of 10.5 km (6.5 miles). Phosphate is ship<strong>pe</strong>d at a weight concentration<br />

of 65–68% by weight. The pump station consists of charging centrifugal pumps and<br />

Worthington plunger pumps. At the terminal station, thickeners and agitated tanks were<br />

installed.<br />

11-7 COPPER SLURRY AND<br />

CONCENTRATE PIPELINES<br />

SLURRY PIPELINES<br />

11.21<br />

Venton and Boss (1996) described the wear of the OK Tedi cop<strong>pe</strong>r concentrate pi<strong>pe</strong>line<br />

in Papua New Guinea. This pi<strong>pe</strong>line is 155 km (95 mi) long. Severe localized wear<br />

caused replacement of certain section of the pi<strong>pe</strong>line. The pi<strong>pe</strong>line was commissioned in<br />

1987. The mine is 156 km (96 mi) from the seaport of Kiunga. About 96 km (60 mi) of<br />

the pi<strong>pe</strong>line uses gravity flow. A booster station was installed to promote flow over the<br />

remaining 60 km (37.5 mi). The normal flow rate was in the range of 85–88 m 3 /hr<br />

(374–388 US gpm) with a weight concentration of 55–60%. The s<strong>pe</strong>ed of flow was in the<br />

range of 1.22–1.4 m/s (4–4.6 ft/sec). The wall thickness was in the range of 5.6–11 mm<br />

(0.22–0.433 in) (Table 11-4). The pi<strong>pe</strong>line o<strong>pe</strong>rated in batches of water and concentrate.<br />

The water was not neutralized by an oxygen scavenger.<br />

Venton and Boss (1996) indicated that an initial pi<strong>pe</strong>line failure occurred in 1991.<br />

They attributed this failure to accentuated wear due to coarser particles, not the design<br />

of the valves. The most severe wear occurred at the change of pi<strong>pe</strong> thickness from<br />

5.6–6.4 mm (0.22–0.25 in). Corrosion was also a factor as no oxygen scavenger had<br />

been used. The o<strong>pe</strong>rators installed corrosion-meter probes in 1992 on the top and bottom<br />

section of the pi<strong>pe</strong>line to monitor wear as loss of wall thickness. Wear of 0.37<br />

mm/year (0.0145 in/year) was measured for the bottom section of the pi<strong>pe</strong>, and wear<br />

of 0.18 mm/year (0.007 in/year) for the top section with continuous <strong>slurry</strong> water/batching.<br />

Venton and Boss (1996) reported that there were four batches of slurries at 500 m 3<br />

(17,657 ft 3 ) and four batches of water of 40–50 m 3 (1,413–1,766 ft 3 ) <strong>pe</strong>r day. To mitigate<br />

against wear, the top size (+106 microns) was cut down to 1%, water batching was eliminated<br />

altogether, and the pi<strong>pe</strong>line was allowed to shut down and restart with <strong>slurry</strong>. Unfortunately,<br />

60 km (37.5 mi) of pi<strong>pe</strong> had to be replaced with thicker walled pi<strong>pe</strong> to continue<br />

o<strong>pe</strong>ration over its anticipated life of 15 years.<br />

TABLE 11-4 Velocity of Flow of Cop<strong>pe</strong>r Concentrate Pi<strong>pe</strong>lines<br />

Pi<strong>pe</strong>line Nominal flow rate Nominal velocity<br />

OK Tedi DN150 (6 in OD) 85 m3 /h 374 USGPM 1.24 m/s 4.07 ft/s<br />

Bonguinville DN 150 (6 in OD) 65–109 227–479 1.23–2.04 4.03–6.7<br />

Freeport DN 100 (4 in OD) 1.1–1.6 3.6–5.25<br />

DN 125 (5 in OD) 1.2–1.5 3.95–4.1<br />

After Venton and Boss (1996).


11.22 CHAPTER ELEVEN<br />

Venton and Boss (1996) described in great detail the o<strong>pe</strong>rating problems of the OK<br />

Tedi pi<strong>pe</strong>line. Their recommendations for better o<strong>pe</strong>ration included the following:<br />

� Installing tracking modules on each pig for pigging the pi<strong>pe</strong>line.<br />

� Replacing the Rockwell Nordstrom valves, which often fail due to inadequate lubrication<br />

in remote valve stations, by other valves. Tests were run on Audco full bore<br />

valves, Mogas metal-seated valves, and Larox high-pressure hydraulically activated<br />

punch valves. The Mogas valves did not require lubrication and lasted 140 o<strong>pe</strong>rating<br />

cycles whereas the Rockwell Nordstrom plug lasted 35 cycles, the Audco ball lasted<br />

20 cycles, and the Larox pinch valve lasted 45 cycles. The Mogas valve was therefore<br />

the most appropriate for this cop<strong>pe</strong>r concentrate pi<strong>pe</strong>line.<br />

One of the largest cop<strong>pe</strong>r mines in the world is o<strong>pe</strong>rated by Minera Escondida Ltd. in<br />

Chile. The mine is located at an altitude of 3100 m (10,170 ft) above sea level. To transport<br />

the cop<strong>pe</strong>r concentrate a pi<strong>pe</strong>line was constructed. The pi<strong>pe</strong>line uses a single pump<br />

station at the beginning of the pi<strong>pe</strong>line and gravity throughout the remainder. The pi<strong>pe</strong>line<br />

spans 165 km (103 mi) of mountainous terrain and transports the <strong>slurry</strong> at a cost of 1–1.5<br />

dollars <strong>pe</strong>r metric ton. This style of transportation is considerably chea<strong>pe</strong>r than the alternative<br />

option of trucks and railroads. Nordstrom valves were used on the pi<strong>pe</strong>line (Boggan<br />

and Buckwalter, 1996).<br />

Bajo Alumbrera is located in northwest Argentina near Catamarca. The plant processes<br />

90,000 tons of ore a day. Cop<strong>pe</strong>r concentrate is ship<strong>pe</strong>d to a port via a 152 mm (6 in)<br />

pi<strong>pe</strong>line over a distance of 320 km. Geho positive displacement diaphragm pumps in the<br />

main pi<strong>pe</strong>line and a couple of booster stations provide the power to pump over such a<br />

long distance.<br />

11-8 CLAY AND DRILLING MUDS<br />

Sellgrem et al. (2000) conducted tests on sand as well as sand–clay mixtures pum<strong>pe</strong>d by<br />

centrifugal pumps. The phosphate clays had a diameter d 50 between 1 �m and 50 �m.<br />

The sands were much coarser with d 50 of 0.64 mm (0.025 in), 1.27 mm (0.05 in), and 2.2<br />

mm (0.09 in). The presence of clay and other particles finer than 75 �m and a concentration<br />

smaller than 20% had a beneficial effect by reducing the head loss and efficiency derating<br />

factor. The data recorded by Sellgrem et al. (2000) should not be applied to a higher<br />

concentration of clays because the viscosity effect introduces a new component to the<br />

equation.<br />

Drilling muds and bentonite are pum<strong>pe</strong>d at high concentration in the oil industry using<br />

positive displacement pumps. Certain important minerals such as bauxite for the aluminum<br />

industry are found in bentonite.<br />

Soft high clay is present in certain cop<strong>pe</strong>r ores. In a diluted form at weight concentration<br />

smaller than 40%, it can be handled fairly well. However, particular attention must be<br />

paid in milling circuits when the concentration may approach 50%, as the viscosity affects<br />

flow and recirculation loads.<br />

Codelco in Chile is one of the largest producers of cop<strong>pe</strong>r in the world. To dispose<br />

tailings to the Ovejeria Tailings Dam, <strong>slurry</strong> had to be pi<strong>pe</strong>d from an elevation of 4000 m<br />

above sea level down to 700 m via a 57 km (36 mile) long pi<strong>pe</strong>line. At the dam, the coarse<br />

and fine are separated using a cyclone station. Three Wirth positive displacement pumps<br />

are then used to pump the coarse material around a 4 km loop at a flow rate of 140 m 3 /hr<br />

(616 US gpm) and at a pressure of 4.0 MPa (580 psi) (Figure 11-11).


11-9 OIL SANDS<br />

SLURRY PIPELINES<br />

11.23<br />

FIGURE 11-11 Tailings solids segragation station and pumping facility for cop<strong>pe</strong>r tailings<br />

at Codelco Andina, Chile, featuring the use of diaphragm pumps. (Courtesy of Wirth Pumps,<br />

Germany.)<br />

In 2001, Canadian oil sands were the most pum<strong>pe</strong>d slurries in the world. Due to this large<br />

amount of <strong>slurry</strong> (1.9 m 3 /s) (ranging up to 30,000 US gpm), pump manufacturers develo<strong>pe</strong>d<br />

new technologies, including technology for froth treatment.<br />

At the Summit Meeting at Quebec during April 2001, Canada encouraged the United<br />

States to invite U.S. corporations to invest billions of dollars in the oil sand fields of Alberta.<br />

Even without new U.S. investments, an estimated 20 billion will be invested between<br />

2000 and 2020. This is a continuously growing industry that will require sophisticated<br />

<strong>slurry</strong> <strong>systems</strong>.<br />

The process of extracting oil from sand is a vast topic and only a few as<strong>pe</strong>cts will be<br />

touched on in this chapter. In basic terms, Alberta, Canada sits on layers of tar-rich sands.<br />

The shallowest layers, which are most accessible for o<strong>pe</strong>n pit mining, are in the north of<br />

Alberta near the Athabasca River and Fort McMurray. Between the first discoveries in the<br />

1930s and the end of the century, a number of technologies were develo<strong>pe</strong>d to the extract<br />

oil from the sand. The initial approach was to heat slurries of tar sand to a tem<strong>pe</strong>rature<br />

that reduced the viscosity of the oil and separated it from the sand. Other technologies develo<strong>pe</strong>d<br />

coannular flows that separated the oil from the sand by degradation of the natural<br />

lumps of oil and sand. More recently, solvents were develo<strong>pe</strong>d that dissolve the tar or oils<br />

out of the sand. The latest solvent-based technologies use lower tem<strong>pe</strong>ratures, reducing<br />

energy costs.<br />

Outside the Athabasca region, the oil sands are located in dee<strong>pe</strong>r layers. The proposed<br />

extraction method pumps hot steam down approximately 100 m (300 ft) of pi<strong>pe</strong> to the oil<br />

sand bed. The steam would then resurface carrying the oil and tar. This technology was


11.24 CHAPTER ELEVEN<br />

originally proposed to recover oil from oil shale in Colorado, U.S., and Queensland, Australia.<br />

Syncrude and Suncor-Muskeg River Shell in conjunction with a Canadian Research<br />

Institute, the Saskachewan Research Institute, the University of Alberta, and the University<br />

of Toronto, develo<strong>pe</strong>d this new technology for oil sand slurries pumping. The concept<br />

of stratified two-layer flows was extensively investigated by these companies and institutes<br />

to handle 63 mm (2.5 in) lumps. By 1999, the price of synthetic oil produced from<br />

processed tar sand by Syncrude and Suncor drop<strong>pe</strong>d low enough to com<strong>pe</strong>te with natural<br />

oil from Texas and the Middle East. A dedicated pi<strong>pe</strong>line from Edmonton, Canada to<br />

Chicago, U.S.A. became the longest pi<strong>pe</strong>line for synthetic fuel.<br />

In a recent pa<strong>pe</strong>r, Sanders et al. (2000) discussed the effects of bitumen on sand hydrotransport<br />

and conducted tests on a number of grades of oil sands. They reported that<br />

pi<strong>pe</strong>line pressure losses due to friction at cold or warm tem<strong>pe</strong>ratures increased with the<br />

length of the pi<strong>pe</strong>. A time de<strong>pe</strong>ndency develo<strong>pe</strong>d, which was attributed to the formation<br />

of a thin coating of bitumen at the wall of the pi<strong>pe</strong>. They defined an equivalent pi<strong>pe</strong><br />

roughness in the presence of bitumen of 650–1150 �m, which is much higher than normal<br />

steel at a roughness of 63 �m. In a 250 mm (10 in) pi<strong>pe</strong>, the presence of fines in the<br />

oil sand <strong>slurry</strong> reduced the deposition velocity to 1.1 m/s, whereas the absence of fines increased<br />

the deposition velocity to 2.7 m/s. Due to the change in s<strong>pe</strong>eds, different approaches<br />

are used in the designs of pi<strong>pe</strong>lines for coarse grade and fine grade ores. The<br />

lower-grade ores, with less bitumen, do not necessarily exhibit this phenomenon of wall<br />

coating and therefore require higher pressure for pumping.<br />

Another pi<strong>pe</strong> coating focused on in numerous tests is the tar coating of pi<strong>pe</strong>s in froth<br />

treatment plants. Under certain conditions, the injection of water through a ring just a few<br />

diameters before the pump suction reduces power consumption and improves the efficiency<br />

of pumps. It is not known whether tar deposits on the im<strong>pe</strong>ller end causes a degradation<br />

of pump <strong>pe</strong>rformance. The physical pro<strong>pe</strong>rties of oils in oil sand would defy any<br />

designer of centrifugal <strong>slurry</strong> pumps and there are no standard methods to account for derating<br />

of <strong>pe</strong>rformance.<br />

McKibben et al (2000) conducted tests on water–oil mixtures (without sand) with<br />

crude oils of a viscosity of 5300–11,200 Pa·s. They found evidence against two popular<br />

theories. First, they showed how the injection water did not form a layer at the wall to reduce<br />

pressure losses as was commonly thought. Instead, it formed slug around the oil and<br />

transported at lower pressure losses. Secondly, they also indicated that the viscosity of the<br />

oil was of no consequence. Therefore, it must be said that the flow of oil, sand, and water<br />

as a mixture is fairly complex.<br />

The Canadian oil sand projects have encouraged the manufacturers of <strong>slurry</strong> pumps to<br />

develop s<strong>pe</strong>cial mechanical seals for <strong>slurry</strong> pumps (Swamanathan et al., 1990).<br />

11-10 BACKFILL PIPELINES<br />

A backfill is essentially a mine residue or tailing pum<strong>pe</strong>d back to fill excavated or mined<br />

pits. A backfill can be mixed with other low-<strong>pe</strong>rmeability materials such as clays to help<br />

seal the area. Particularly in the case of underground backfilling, the water content should<br />

be minimized to avoid costly dewatering. Backfill slurries are therefore dense, with a high<br />

weight concentration (around 50–65%). Multistage centrifugal pumps or positive displacement<br />

pumps are used to transport them (Figure 11-12).<br />

Steward (1996) conducted an empirical study of vertical and horizontal pi<strong>pe</strong>lines. He<br />

demonstrated how backfill consisting of fine and coarse material could be classified


through cyclones in order to separate them. The cyclone underflow was then drained by<br />

gravity. In the South African gold and uranium mines, Steward (1996) reported flow<br />

s<strong>pe</strong>eds as high as 12 m/s (40 ft/s). These extremely high s<strong>pe</strong>eds are the cause of rapid<br />

wear and erosion–corrosion.<br />

To support and reinforce an underground excavated area after the ore has been removed,<br />

the backfill (including both fine and coarse material) is thickened. The product is<br />

called full plant tailings. Steward (1996) defined the particle sharpness as the rate of directional<br />

change in the particle <strong>pe</strong>rimeter.<br />

Pi<strong>pe</strong> wear is measured as the rate of mass loss <strong>pe</strong>r unit of time (kg/s or slugs/s). Wear<br />

in a pi<strong>pe</strong>line is an exponential function of the flow s<strong>pe</strong>ed:<br />

dw/dt = KV n<br />

Since it is also a function of other parameters, Steward (1996) proposed the following<br />

function:<br />

where<br />

f = function of<br />

S m = s<strong>pe</strong>cific gravity of mixture<br />

SLURRY PIPELINES<br />

dw/dt = f (S m, V, D I, d 90, SI, M)<br />

11.25<br />

FIGURE 11-12 Backfilling of very dense <strong>slurry</strong> using diaphragm pumps. (Courtesy of Wirth<br />

Pumps, Germany.)


11.26 CHAPTER ELEVEN<br />

V = velocity of flow<br />

D I = pi<strong>pe</strong> inner diameter<br />

d 90 = 90% passing diameter of particles<br />

SI = sharpness index<br />

M = mass of solids <strong>pe</strong>r unit length of pi<strong>pe</strong><br />

From his tests, Steward (1996) derived the following empirical equation (in SI units)<br />

(Table 11-5):<br />

where<br />

m 1 = –6.31374<br />

m 2 = 0.3186193<br />

m 3 = –0.131869<br />

m 4 = 0.0054758<br />

m 5 = 1.7709578<br />

m 6 = 0.6162088<br />

m 7 = 6.5961888<br />

log 10(dw/dt) = m 1S m + m 2V + m 3D I + m 4d 90 + m 5SI + m 6M + M 7<br />

Coetzee (1990) determined that one-third of the loss of pi<strong>pe</strong>line wall thickness associated<br />

with pumping mine water is due to corrosion because mine water is often acidic.<br />

Since corrosion is an important contributor to wear of backfilled pi<strong>pe</strong>s, it became evident<br />

that lining the pi<strong>pe</strong>s was necessary. To compare piping materials, tests were conducted by<br />

Steward (1996) and indicated that a polyurethane rubber at a Shore hardness 55 Shore A<br />

provided the best pi<strong>pe</strong>line protection in a test with <strong>slurry</strong> pum<strong>pe</strong>d at a s<strong>pe</strong>ed of 3 m/s. By<br />

comparison, ASTM steel 106 grade B wore seven times faster than polyurethane 82 Shore<br />

A, or high-density polyethylene.<br />

Steward (1996) indicated that the mixing of cementitious binders with <strong>slurry</strong> could reduce<br />

wear considerably in backfill applications.<br />

Backfill paste is formed by dewatering <strong>slurry</strong> of tailings (thickening and filtering).<br />

Mixing dewatered <strong>slurry</strong> with cement (3%–5%) produces a stiff backfill (1.5–3.5 MPa<br />

strength, or 218–508 psi). Coarse aggregates (


11-11 URANIUM TAILINGS<br />

Uranium is considered an important material for nuclear energy production. The pumping<br />

of uranium slurries is more complicated than pumping most slurries due to its radioactivity.<br />

As a result, uranium tailings disposal <strong>systems</strong> should involve a health s<strong>pe</strong>cialist and an<br />

environmental engineer.<br />

Uranium ores vary dramatically from one site to another. In certain areas (for example,<br />

Saskatchewan, Canada, Wyoming, U.S.A., and South Australia, Australia), the ore is<br />

rich. In other areas (for example: the mineral sands of Egypt, or the South African<br />

gold/uranium mines), uranium is essentially a by-product.<br />

Due to radioactivity, the design of the tailings disposal system must be left to a lining<br />

s<strong>pe</strong>cialist. It is extremely important that no seepage be released. Four main ty<strong>pe</strong>s of liners<br />

are usually selected:<br />

� Geological liners<br />

� Clay liners<br />

� Synthetic liners such as synthetic membranes, gunite, asphalitic concrete, or sprays<br />

� Compacted soil or soil cement<br />

Some of the slimes from the tailings can be used to form a lining later. Clays, es<strong>pe</strong>cially<br />

bentonites, are very good liners with a <strong>pe</strong>rmeability as low as 10 –7 cm/s (4 × 10 –8 in/sec).<br />

To dissolve uranium in milling, two processes are used:<br />

1. Acid leaching<br />

2. Alkaline leaching<br />

SLURRY PIPELINES<br />

In acid leaching, sulphuric acid is used in a complex ion-exchange or solvent extraction<br />

process to produce yellowcake of very high purity. Various metals (such as vanadium,<br />

arsenic, nickel, iron, cop<strong>pe</strong>r, etc.) may be leached in this process. Chemicals involved<br />

in this process include sulphuric acid, ammonium nitrate, sodium chloride, amines, alcohols,<br />

kerosene, and ammonia. Considerable process water has to be derived from reclaim<br />

water of the tailings and returned to the mine for preparing the <strong>slurry</strong>.<br />

The alkaline process is less common than the acid process and provides lower recovery<br />

rates of uranium. For both processes, the ore must be first reduced to a size smaller than 75<br />

�m (mesh 200). Important radioactive pollutants from either process are 226Ra (radium<br />

226) and 222Rn (radon 222). By adding barium chloride in settling ponds or lagoons, radium<br />

is treated. Radium-containing slurries and sludges are the subject of extensive research.<br />

Radium 226 decays to form radon gas from the tailings and is dangerous to health.<br />

The pH of solutions from the solvent-extraction process is typically low, generally<br />

1–2. Conventional carbon steel pi<strong>pe</strong>s may be inferior to high-density polyethylene pi<strong>pe</strong>s<br />

in resisting combined erosion and corrosion.<br />

11-12 CODES AND STANDARDS FOR<br />

SLURRY PIPELINES<br />

11.27<br />

Safety is of paramount importance with some of the modern pi<strong>pe</strong>lines. Slurry is heavier<br />

than water, and the rupture of a pi<strong>pe</strong>line can have disastrous effects on the environment


11.28 CHAPTER ELEVEN<br />

and may cause accidental death. For this reason the engineer and contractor should follow<br />

certain codes and standards. The ASME B31.11 is basically the only standard s<strong>pe</strong>cific<br />

to <strong>slurry</strong> pi<strong>pe</strong>lines. However, the American Petroleum Institute has many useful<br />

guidelines.<br />

The ASME (American Society of Mechanical Engineers) develo<strong>pe</strong>d in 1989 a s<strong>pe</strong>cial<br />

section to its piping code B31 called ASME B31.11-1989—Slurry Transportation Piping<br />

Systems. This standard provides useful guidelines for o<strong>pe</strong>ration and maintenance of <strong>slurry</strong><br />

pi<strong>pe</strong>lines and piping <strong>systems</strong>. Because the code came into existence at the end of the<br />

1980s, it is still not well known. Consultant engineers need to refer to it, particularly in<br />

the case of all-metal piping.<br />

It is the ultimate responsibility of the engineer and the contractor to provide safe procedures<br />

for the o<strong>pe</strong>ration and maintenance of the <strong>slurry</strong> piping system. The code provides<br />

some appropriate guidelines. The code requires an established plan for o<strong>pe</strong>ration and a<br />

maintenance procedure. The plan must be written with instructions to o<strong>pe</strong>rate and maintain<br />

the pi<strong>pe</strong>line. It must include provisions for control of external as well as internal corrosion<br />

and erosion of new and existing pi<strong>pe</strong>s. It must describe emergency procedures to<br />

follow in case of system failures, accidents, and other hazards. Employees need to be<br />

trained in these emergency procedures. Because pi<strong>pe</strong>lines eventually cross rural or urban<br />

areas and could have an impact on the environment, the local authorities should be consulted<br />

when establishing such a plan.<br />

The plan must detail procedures to review and reflect on important changes in the conditions<br />

affecting the safety and integrity of the piping <strong>systems</strong>. A methodology must be in<br />

place to patrol and report on changes in construction activities, railroad and highway<br />

crossings, and urban and commercial activity. Obviously, no one should come with an excavator<br />

and damage a buried pi<strong>pe</strong>line because he did not know that it existed. An abandonment<br />

plan must include procedures for shutting down and abandoning a pi<strong>pe</strong>line, and<br />

must include appropriate cleanup procedures before abandonment.<br />

The o<strong>pe</strong>rating pressure includes the steady-state pressure to overcome friction as well<br />

as the static head. The code requires that the level of pressure rise due to surges not exceed<br />

at any point the internal design pressure by more than 10%.<br />

Markers showing the existence of the pi<strong>pe</strong>line must be installed on each side of a public<br />

road, railroad crossing, or crossing of navigable water. Markers are not required for<br />

off-shore pi<strong>pe</strong>lines such as those used in subsea tailings disposal. The following standard<br />

should be consulted: API RP1109: Marking Liquid Petroleum Pi<strong>pe</strong>line Faculties 1993.<br />

This code provides guidelines for the <strong>pe</strong>rmanent marking of liquid <strong>pe</strong>troleum pi<strong>pe</strong>line<br />

transportation facilities. It covers the design, message writing, installation, placement, ins<strong>pe</strong>ction,<br />

and maintenance of markers and signs for inland waterway crossings and onshore<br />

crossing. It is a short standard of 12 pages that the consultant engineer should modify<br />

to suit the <strong>slurry</strong> pi<strong>pe</strong>line.<br />

Patrolling pi<strong>pe</strong>lines on a <strong>pe</strong>riodic basis is recommended, particularly to ensure that all<br />

valves remain accessible (and are not lost between growing vegetation), to ensure that<br />

emergency diversion ditches remain free of debris, grass, and leaves. The code recommends<br />

limiting patrols to a maximum of one <strong>pe</strong>r month. Underwater crossing should be<br />

ins<strong>pe</strong>cted particularly after a flood or a heavy storm that may have caused mechanical<br />

damage.<br />

Pi<strong>pe</strong>line repairs need to be <strong>pe</strong>rformed by qualified <strong>pe</strong>rsonnel. ASME B31.11 recommends<br />

using the following publication of the American Petroleum Institute as a guideline:<br />

API RP1111: Design, Construction, O<strong>pe</strong>ration and Maintenance of Off-shore Hydrocarbon<br />

Pi<strong>pe</strong>line and Risers (Third Edition 1999).<br />

Safety guidelines for a pi<strong>pe</strong>line crossing are covered by the following document: API<br />

RP1102: Steel Pi<strong>pe</strong>lines Crossing Railroads and Highways (Sixth Edition, April 1993).


SLURRY PIPELINES<br />

11.29<br />

API RP1102 covers design, installation, and testing to ensure safe crossing of steel<br />

pi<strong>pe</strong>lines under railroads and highways.<br />

High-pressure <strong>slurry</strong> pi<strong>pe</strong>lines, particularly some of the newer 2500 psi pi<strong>pe</strong>lines, must<br />

be safely welded. Some of these use API Grade 5LX65 or 5LX70 steel. It is recommended<br />

to implement API Standard 1104: Welding of Pi<strong>pe</strong>lines and Related Facilities (Nineteenth<br />

Edition, 1999.) This standard was obviously written for the oil and gas industry. It covers<br />

gas and arc welding to complete high-quality welds of carbon steel as well as low-alloy<br />

steel piping. It covers different ty<strong>pe</strong>s of welding processes, such as shielded metal arc welding,<br />

submerged arc welding, gas–tungsten arc welding, gas–metal arc welding, as well<br />

methods to test and ins<strong>pe</strong>ct welds by radiography. When a pi<strong>pe</strong>line is constructed, a standard<br />

procedure is to conduct hydrostatic tests. The American Petroleum Institute develo<strong>pe</strong>d<br />

the following recommended practice document: RP1110: Pressure Testing of Liquid Petroleum<br />

Pi<strong>pe</strong>line (Fourth Edition, March 1997). This document is of interest to the <strong>slurry</strong><br />

pi<strong>pe</strong>line engineer because it descries the minimum procedures to be followed as well as<br />

some of the equipment used during hydrostatic testing. Because <strong>slurry</strong> pi<strong>pe</strong>lines are now<br />

built in lengths from few kilometers (or few miles) to hundreds of kilometers (or miles) and<br />

are often unpatroled and buried in remote and isolated regions, control engineers have develo<strong>pe</strong>d<br />

methods to monitor pi<strong>pe</strong>lines for pressure or leakage. One such system is called<br />

SCADA (system control and data acquisition). The American Petroleum Institute has develo<strong>pe</strong>d<br />

a number of useful documents to consult. Publication 1113: Developing a Pi<strong>pe</strong>line<br />

Su<strong>pe</strong>rvisory Control Center (Third Edition, February 2000) presents six lists of general<br />

considerations to design a center to monitor and control a pi<strong>pe</strong>line.<br />

With the advent of modern computer-based system for monitoring pi<strong>pe</strong>lines, API develo<strong>pe</strong>d<br />

an appropriate standard: Std 1130: Computational Pi<strong>pe</strong>line Monitoring (First<br />

Edition, October 1995). This standard is particularly important in detecting anomalies,<br />

which can be attributed to leaks, rupture of the line, etc. It covers algorithmic monitoring<br />

tools to help the <strong>pe</strong>rson in charge of control and monitoring detect such anomalies.<br />

With the importance accorded to early detection of pi<strong>pe</strong>line leaks, the American Petroleum<br />

Institute formed a taskforce in 1989 with the University of Idaho to evaluate commercial<br />

software for leak detection. A report was published as Publ. 1149: Pi<strong>pe</strong>line Variable<br />

Uncertainties and Their Effects on Leak Delectability (First Edition, 1993). This<br />

document was followed by a second and rather useful publication, Publ. 1155: Evaluation<br />

Methodology for Software-Based Leak Detection Systems (First Edition, 1995). These<br />

publications and standards should not be imposed by the consultant engineer without reference<br />

notes and modifications as related to <strong>slurry</strong>. They have been listed as useful tools<br />

because most high-pressure pi<strong>pe</strong>lines are transporting oil and gas and these industries<br />

have contributed greatly to the safety of <strong>systems</strong> similar to the ones the <strong>slurry</strong> engineer is<br />

designing. The ASME Code B31.11 recommends certain <strong>pe</strong>rmanent repair procedures<br />

when particular problems have been detected. When hoop stress is detected exceeding<br />

20% above the minimum yield stress, followed by a reduction of thickeners, the code recommends<br />

making the following <strong>pe</strong>rmanent repairs:<br />

� Remove all gouges and grooves having a depth in excess of 12.5% of the nominal pi<strong>pe</strong><br />

thickness.<br />

� Remove dents that affect the pi<strong>pe</strong> curvature at the pi<strong>pe</strong> seam or at any girth weld.<br />

� Remove dents that exceed 7.5% of the nominal pi<strong>pe</strong> diameter.<br />

� Remove or repair all arc burns.<br />

� Remove or repair all cracks.<br />

� Remove or repair all welds considered to be im<strong>pe</strong>rfect by the stipulations of the code.


11.30 CHAPTER ELEVEN<br />

� Repair areas where the thickness is smaller than the recommended thickness for pressure<br />

minus the corrosion allowance.<br />

� Remove or repair all sections containing leaks.<br />

Although the code recommends taking out of service defective sections of the pi<strong>pe</strong>line, it<br />

accepts that this is not always possible. It does, therefore, allow for certain methods of repair<br />

while the pi<strong>pe</strong>line is in service, such as the use of mechanically split sleeves, hot tapping,<br />

encirclement of welds, etc. The engineer should, however, be aware that these methods<br />

may not apply to rubber, polyurethane, or HDPE-lined steel pi<strong>pe</strong>s as they would<br />

damage the internal lining by heat.<br />

New replacement sections of a pi<strong>pe</strong>line must be subject to a pressure test after their installations,<br />

when the pi<strong>pe</strong>line is o<strong>pe</strong>rating at a hoop stress of more than 20% of the s<strong>pe</strong>cified<br />

minimum yield strengths.<br />

The control of corrosion and erosion of <strong>slurry</strong> pi<strong>pe</strong>lines is covered by Chapter VIII of<br />

the ASME Code B31.11. The code correctly points out that in certain cases erosion is an<br />

accelerating factor in the internal corrosion of pi<strong>pe</strong>lines, by effectively removing scales,<br />

oxides, films, and lining.<br />

Buried steel pi<strong>pe</strong>s can be subject to corrosion. The National Association of Corrosion<br />

Engineers develo<strong>pe</strong>d the following standard: RP0196-96: Control of External Corrosion<br />

on Underground or Submerged Metallic Piping Systems. This standard provides a<br />

methodology for determining the need for corrosion control, cathodic protection and design,<br />

installation of cathodic protection <strong>systems</strong>, and control of interference currents.<br />

Sacrificial anodes, such as magnesium blocks, are installed in difficult soils such as<br />

those containing chlorides in certain deserts. Due to the difference between magnesium<br />

and steel on the galvanic corrosion chart, the system is designed so that the established<br />

current causes corrosion to the magnesium blocks rather than to the steel pi<strong>pe</strong>. Monitoring<br />

galvanic corrosion is important and the reader may consult the CEA Report 54276, “Cathodic<br />

Protection on Buried Pi<strong>pe</strong>lines,” as well as TPCII, “A Guide to the Organization of<br />

Underground Corrosion Control Coordinating Committees.”<br />

Monitoring cathodic protection is emphasized by ASME B31-11, which recommends<br />

testing at two-month intervals or less of all cathodic protection rectifiers and connected<br />

protective devices. Erosion–corrosion of <strong>slurry</strong> pi<strong>pe</strong>lines may be reduced using s<strong>pe</strong>cial<br />

corrosion inhibitors, by avoiding sharp corners and short elbows, and providing some<br />

lining such as HDPE, rubber, or polyurethane. Monitoring of the pi<strong>pe</strong>line on a yearly<br />

basis and at intervals that do not exceed 15 months is recommended by B31.11. Areas<br />

prone to more rapid localized erosion–corrosion should be monitored more frequently as<br />

dictated by ex<strong>pe</strong>rience. Corrective measures should be established on the basis of historical<br />

leaks.<br />

Records must be kept for location of cathodic protection facilities, repairs of localized<br />

wear problems, and for frequency of such repairs.<br />

11-13 CONCLUSION<br />

Extensive ex<strong>pe</strong>rience has been gained over the years on the pumping of different minerals<br />

and tailings. Some particles are ground to a very fine range and flow as non-Newtonian<br />

mixtures, and some are as coarse as 50 mm (2 in), as in the coal and oil sand industries.<br />

Despite all the promise of coal, only one long coal pi<strong>pe</strong>line has been built in the United<br />

States—the Black Mesa Pi<strong>pe</strong>line. Another long pi<strong>pe</strong>line that transports coal is the<br />

Novo-Siberski pi<strong>pe</strong>line in Siberia, Russia. Other minerals such as cop<strong>pe</strong>r concentrate,


iron oxide, limestone, and phosphate have become a normal part of the engineering of<br />

mines in very remote and isolated regions such as Escondida, Bajo Alumbrera, Antemina,<br />

Samarco, etc. The science of <strong>slurry</strong> pi<strong>pe</strong>line engineering has therefore blossomed into<br />

very practical schemes.<br />

Instrumentation and remote monitoring have made <strong>slurry</strong> pi<strong>pe</strong>lines safer, better protected<br />

against leakage, freezing, and surging.<br />

Great progress has been made with pumping sets. Centrifugal pumps are now available<br />

up to a design pressure of 7 MPa (1000 psi). Positive displacement pumps push the<br />

limit of use to 17.5 MPa (2500 psi). Lockhop<strong>pe</strong>rs are available to pump very coarse solids<br />

(>50 mm or >2 in) up to distances in excess of 105 km (60 mi).<br />

A successful pi<strong>pe</strong>line project de<strong>pe</strong>nds on pro<strong>pe</strong>r economics. This will be the topic of<br />

the next chapter.<br />

11-14 REFERENCES<br />

SLURRY PIPELINES<br />

11.31<br />

Abbot, J. 1965. Use of Hydrocyclones for Thickening and Recovery in the National Coal Board. Filtration<br />

and Separation, 2, 3, 204–208, 234.<br />

Abulnaga, B. E. 1990. An Internal Combustion Engine Featuring the Use of an Oscillating Liquid<br />

Column and a Hydraulic Turbine to Convert the Energy of Fuels. Australian Patent AU-B-<br />

20956/88.<br />

Adams, W. I. 1986. Polyethylene Pi<strong>pe</strong>lines for Slurry Transportation. In 11th International Conference<br />

on Coal Technology. Washington, D.C.: Coal and Slurry Technology Association.<br />

Anand, S., S. K. Ghosh, S. Govindan, and D. B. Nayan. 1986. Maton Rock Phosphate Concentrate<br />

Pi<strong>pe</strong>line. Working pa<strong>pe</strong>r, BHRA Group, Hydrotransport 10, Innsbruck.<br />

Boggan, J. and R. Buckwalter. 1996. Slurry Pi<strong>pe</strong>line Helps Remedy Corrosion at Record Height.<br />

Pi<strong>pe</strong>line and Gas Journal, 223.<br />

Bomberger, D. R. 1965. Hexavalant Chromium Reduces Corrosion in a Coal-Slurry Pi<strong>pe</strong>line. Materials<br />

Protection, 4, 1, 41–48.<br />

Brackebush, F. W. 1994a. Basics of Paste Backfill Systems. Mining Engineering, 46, 1175–1178.<br />

Brackebush, F. W. 1994b. Basics of Paste Backfill Systems. Mining Engineering, 47, 1041–1042.<br />

Brooks, D. A. and C. H. Dodwell. 1985. The Economic and Technical Evaluation of Slurry Pi<strong>pe</strong>line<br />

Transport Techniques in the International Economic Coal Trade. In 10th International Conference<br />

on Coal Technology, Lake Tahoe, Nevada. Washington, D.C.: Coal and Slurry Technology<br />

Association.<br />

Buckwalter, R. and A. Walters. 1989. Selection of coal <strong>slurry</strong> pi<strong>pe</strong>line technologies for gasification<br />

combined power cycle plants. In Proceedings of the 14th International Conference on Coal<br />

and Slurry Technology. Washington, DC: The Coal and Slurry Association.<br />

Burgess, K. E. 2000. Froth Pumping. Technical Bulletin No. 28. Sidney Australia: Warman International.<br />

Coetzee, R. 1990. Wear and Corrosion of Mild Steel Tubes in Backfill and Backfill Feed Filtrate.<br />

Report 332891. Physical Metallurgy Division. Council for Mineral Technology, South Africa.<br />

Ercolani, D., E. Carniani, S. Meli, L. Pelligrini, and M. Primercio. 1988. Shear Degradation of Concentrated<br />

Coal–Water Slurries in Pi<strong>pe</strong>line Flows. In 13th International Conference on Coal<br />

Technology, Denver, CO. Washington, D.C.: Coal and Slurry Technology Association.<br />

Faddick, R. R. 1982. Ship loading Coarse-Coal Slurries. Working pa<strong>pe</strong>r A-3, in 8th International<br />

Conference on Solids in Pi<strong>pe</strong>s, Johannesburg, South Africa.<br />

Gandhi, R. G. 1985. Fosferil phosphate <strong>slurry</strong> pi<strong>pe</strong>line. In 10th International Conference on Coal<br />

Technology, Lake Tahoe, Nevada. Washington, D.C.: Coal and Slurry Technology Association.<br />

Garbett, E. S. and S. M. Yiu. 1988. The Effect of Convective Heat on the Disintegration of a<br />

Coal–Water Mixture in Pneumatic Atomization. In 13th International Conference on Coal<br />

Technology, Denver, CO. Washington, D.C.: Coal and Slurry Technology Association.<br />

Gillies, R. G., J. Schaan, R. J. Sumner, M. J. McKibben, and C. A. Shook. 2000. Deposition Veloci-


11.32 CHAPTER ELEVEN<br />

ties for Newtonian Slurries in Turbulent Flows. Canadian Journal of Chemical Engineering,<br />

78, 4, 704–708.<br />

Hayashi, H. et al. 1980. Some Ex<strong>pe</strong>rimental Studies on Iron Concentrate Slurry Transport in Pilot<br />

Plant. Working pa<strong>pe</strong>r, BHRA Group, Hydrotransport 7, Sendai, Japan.<br />

Hughes, C. V. 1986. Coal Slurry Pump Development Update. Mainline Pumps for the Belovo-<br />

Novosibirsk Pi<strong>pe</strong>line. In 11th International Conference on Coal Technology, Washington,<br />

D.C.: Coal and Slurry Technology Association.<br />

Klose, R. B. and H. W. Mahler. 1982. Investigations into the hydraulic transportation behaviour of<br />

ore and coal sus<strong>pe</strong>nsions with coarse particles. In Hydrotransport 8, Johannnesburg. Cranfield,<br />

UK: BHRA Group.<br />

Kreusing, H., and F. H. Franke. 1979. Investigations on the Flow and Pumping Behavior of Coal–Oil<br />

Mixtures with Particular Reference to the Injection of Coal–Oil Slurry in the Blast Furnace.<br />

Working pa<strong>pe</strong>r C-2, BHRA Group, Hydrotransport 6, BHRA.<br />

Landel, R. F., B. G. Mosen, and A. J. Bauman. 1963. In 4th International Conference on Rheology,<br />

Brown University, Part 3, p. 663. New York: Interscience Publishers.<br />

Leninger, D., W. Erdmann, and R. Kohling. 1978. Dewatering of Hydraulically Delivered Coal.<br />

Working pa<strong>pe</strong>r E-7, BHRA Group, Hydrotransport 5, Hanover.<br />

Lokon, H. B., P. W. Johnson, and R. R. Horsley. 1982. A “Scale-up” Model for Predicting Head<br />

Loss Gradients in Iron Ore Slurry Pi<strong>pe</strong>lines. Working pa<strong>pe</strong>r B-2, BHRA Group, Hydrotransport<br />

8.<br />

Madsen, B. W., S. D. Cramer, and W. K. Collins. 1995. Corrosion in a Phosphate Pi<strong>pe</strong>line. Materials<br />

Performance, 34, 70–73.<br />

McKibben, M., R. G. Gilles, and C. A. Shook. 2000. A Laboratory Investigation of Horizontal Well<br />

Heavy Oil–Water Flows. Canadian Journal of Chemical Engineering, 78, 734–751.<br />

Miller, J. W. and H. L. Hoyt. 1988. Evaluation of Polymers as Sus<strong>pe</strong>nding Aids for Coal–Water<br />

Slurries. In 13th International Conference on Coal Technology. Washington, D.C.: Coal and<br />

Slurry Technology Association.<br />

Morway, A. J. 1965. Stabilized Oiled Coal Slurry in Water. US Patent 31,201,168 assigned to Esso<br />

Research & Engineering Co. N.J., USA.<br />

Nordin, M. 1982. Slurry for Sale. Working pa<strong>pe</strong>r F-2, BHRA Group, Hydrotransport 8.<br />

Olofinsky, E. P. 1988. Belovo-Novosibirsk Coal Transportation Pi<strong>pe</strong>line. In 13th International Conference<br />

on Coal Technology, Denver, CO. Washington, D.C.: Coal and Slurry Technology Association.<br />

Peterson, A. J. C. and K. Mackie. 1996. An Economic and Technical Assessment of the Hydraulic<br />

Transport of Phosphate Ore. BHRA Group, Hydrotransport 13.<br />

Pertuit, P. 1985. Gladstone Limestone Slurry Pi<strong>pe</strong>line. In 10th International Conference on Coal<br />

Technology, Lake Tahoe, Nevada. Washington, D.C.: Coal and Slurry Technology Association.<br />

Pertuit, P. 1985. Goiasferil Phosphate Pi<strong>pe</strong>line. In 10th International Conference on Coal Technology,<br />

Lake Tahoe, Nevada. Washington, D.C.: Coal and Slurry Technology Association.<br />

Pertuit, P. 1985. Hindustan Zinc Phosphate Pi<strong>pe</strong>line. In 10th International Conference on Coal Technology,<br />

Lake Tahoe, Nevada. Washington, D.C.: Coal and Slurry Technology Association.<br />

Pi<strong>pe</strong>lin, A. P., M. Weintraub, and A. A. Orning. 1966. Report of Investigation No. 6743, prepared for<br />

the US Bureau of Mines.<br />

Sanders, R. S., A. L. Ferre, W. B. Maciejewski, R. Giles, and C. Shook. 2000. Bitumen Effects on<br />

Pi<strong>pe</strong>line Hydraulics during Oil–Sand Hydrotransport. Canadian Journal of Chemical Engineering,<br />

78, 4, 731–742.<br />

Schaan, J., R. J. Sumner, R. G. Gillies, and C. A. Shook. 2000. The Effect of Particle Sha<strong>pe</strong> on<br />

Pi<strong>pe</strong>line Friction for Newtonian Slurries of Fine Particles. Canadian Journal of Chemical Engineering,<br />

74, 4, 717–725.<br />

Sellgrem, A., G. Addie, and S. Scott. 2000. The Effect of Sand–Clay Slurries on the Performance of<br />

Centrifugal Pumps. Canadian Journal of Chemical Engineering, 78, 4, 764–769.<br />

Shook, C. A., D. B. Haas, W. H. W. Husband, and M. Smail. 1979. Degradation of Coarse Coal Particles<br />

during Hydraulic Transport. Working pa<strong>pe</strong>r C-1, BHRA Group, Hydrotransport 6.<br />

Steward, N. R. 1991. The Determination of Wear Relationships for FORSOC Fillset Binder Modified<br />

Classified Tailings at High Relative Density. Report for Gold and Uranium Division of the<br />

Anglo American Corporation of South Africa Ltd.


SLURRY PIPELINES<br />

11.33<br />

Swamanthan, S., A. Fair, and J. Wong. 1990. In Search of Mechanical Seals for Slurry Pumps. In<br />

15th International Conference on Coal Technology, Lake Tahoe, Nevada. Washington, D.C.:<br />

Coal and Slurry Technology Association.<br />

Steward, N. B. 1996. An Empirical Evaluation of the Wear of Backfill Transport Pi<strong>pe</strong>lines. Working<br />

pa<strong>pe</strong>r, BHRA Group, Hydrotransport 13, Cranfield, England.<br />

Thomas, A. D. 1976. Scale-up Methods for Pi<strong>pe</strong>line Transport of Slurries. Int. Journal of Mineral<br />

Processing, 3, 51–69.<br />

Thompson, T. L. 1985. La Perla/Hercules Pi<strong>pe</strong>line. In 10th International Conference on Coal Technology,<br />

Lake Tahoe, Nevada. Washington, D.C.: Coal and Slurry Technology Association.<br />

Tian, H., G. Addie, and R. S. Hagler. 1996. Development of Corrosion Resistant White Irons for Use<br />

in Phos-acid Service. Pa<strong>pe</strong>r presented at the Annual Conference of Central Florida section of<br />

the American Institute of Chemical Engineers.<br />

Tillotson, I. S. 1953. Hydraulic Transportation of Solids. M.n. Congress Journal, 39, 1, 41–44.<br />

Vanderpan, R. I. 1982. Pro<strong>pe</strong>r Pump Selection for Coal Preparation Plants. In World Coal. San Francisco:<br />

Miller Freeman Publications.<br />

Venton, P. B. 1982. The Gladstone Pi<strong>pe</strong>line. Working pa<strong>pe</strong>r A-4, BHRA Group, Hydrotransport 8.<br />

Venton, P. B. and T. J. Boss. 1996. An Analysis of Wear Mechanisms in the 155 km OK Tedi Cop<strong>pe</strong>r<br />

Concentrate Slurry Pi<strong>pe</strong>line. Working pa<strong>pe</strong>r, BHR Group, Hydrotransport, 533–548.<br />

Walker, C. I. 1993. A New Alloy for Phosphoric Acid Slurries. Pa<strong>pe</strong>r presented at the 1993 Clearwater<br />

Convention, American Institute of Chemical Engineers.<br />

Weston, M. D. 1985. SAMARCO Pi<strong>pe</strong>line. In 10th International Conference on Coal Technology,<br />

Lake Tahoe, Nevada. Washington, D.C.: Coal and Slurry Technology Association.<br />

Weston, M. D. and L. Worthen. 1987. Chevron Phosphate Slurry Pi<strong>pe</strong>line commissioning and startup.<br />

In Proceedings of the 12th International Conference on Coal and Slurry Technology.<br />

Washington, DC: The Coal and Slurry Association.<br />

Editorial Articles<br />

Moving Mountains through a Slurry Pi<strong>pe</strong>line. Engineering and Mining Journal, 195, 94–95, 1994.<br />

A Conductance Based Solids Concentration Sensor for Large Diameter Slurry Pi<strong>pe</strong>lines. Journal of<br />

Fluid Engineering, 122, 4<br />

Variety of Slurry Pumps in Taconite Processing Plants. Skillings Mining Review, 70, 32, Aug 8,<br />

1981.<br />

Further Readings<br />

Abulnaga, B. A. 2000. A Review of the Yichang Phophate Pi<strong>pe</strong>line Feasibility Study. HATCH, unpublished.<br />

USSR Plans Coal Slurry Pi<strong>pe</strong>lines. Oil and Gas Journal, 82, 58–59, 1984.<br />

Braca, R. M. 1988. Use Needs Coal Slurry Pi<strong>pe</strong>line. Pi<strong>pe</strong>line and Gas Journal, 215, 32–36.<br />

Catalano, L. 1983. Railroads Kill Eminent Domain for Coal-Slurry Pi<strong>pe</strong>lines. Power Journal, 127, 9.<br />

Harvey, W. W. and Hossain, M. A. 1987. Co-recovery of Chromium from Domestic Nickel Laterites.<br />

Journal of Metals, 39, 21–25.<br />

Mahr, D. and B. Robert. 1986. Coal Slurry Pi<strong>pe</strong>lines Overland Belt Conveyors See Bright Future.<br />

Power Engineering, 90, 24–28.<br />

Maki, G. A. and D. M. Smith. 1983. Potash Mines and Mining/Saskatchewan/Thickeners/Design.<br />

CIM Bulletin, 76, 57–62.<br />

Maki, G. A., R. G. Roden, and P. J. Fullman. 1990. Stacking of Potash Mill Tailings. CIM Bulletin,<br />

83, 96–98.<br />

Marrey, D. T. 1985. Exporting Colorado Water in Coal Slurry Pi<strong>pe</strong>line. Journal of Water Resources<br />

Planning and Management, 111, 207–221.<br />

Nalziger, R. H. 1988. Ferrochromium from Domestic Lateritic Chromites. Journal of Metals, 40,<br />

34–37.<br />

Nasr-El-Din, H., C. A. Shook, and M. N. Esmail. 1984. Isokinetic Probe Sampling from Slurry<br />

Pi<strong>pe</strong>lines. Canadian Journal of Chemical Engineering, 62, 179–185.<br />

Postlethwaite, J. 1987. The Control of Erosion–Corrosion in Slurry Pi<strong>pe</strong>lines. Materials Performance,<br />

26, 41–45.<br />

Postlethwaite, J., M. H. Dobbin, and K. Bergevin. 1986. The Role of Oxygen Mass Transfer in the<br />

Erosion-Corrosion of Slurry Pi<strong>pe</strong>lines. Corrosion, 42, 514–521.


11.34 CHAPTER ELEVEN<br />

Schaan, J., and C. A. Shook. 2000. Anomalous Friction in Slurry Flows. Canadian Journal of Chemical<br />

Engineering, 78, 4, 726–730.<br />

Shvartsburd, V. 1983. Pi<strong>pe</strong>lining and Burning Coal, Here are Important Criteria for Designing Coal<br />

Slurry Pi<strong>pe</strong>lines. Oil and Gas Journal, 81, 91–95.<br />

Wasp, E. J. 1983. Slurry Pi<strong>pe</strong>lines. Scientific American, 249, 48–55.


CHAPTER 9<br />

POSITIVE DISPLACEMENT<br />

PUMPS<br />

9-0 INTRODUCTION<br />

Positive displacement <strong>slurry</strong> pumps and mud transfer pumps play a major role in a number<br />

of industries such as mining and metallurgical processes, chemicals, power generation,<br />

porcelain and ceramics, and sugar refining. These pumps are versatile, efficient,<br />

and suitable for pressures up to 17.3 MPa (2500 psi). Positive displacement pumps<br />

have gained acceptance on long-distance mineral concentrate pi<strong>pe</strong>lines as their high<br />

capital cost is recu<strong>pe</strong>rated through lower installation cost of electric <strong>systems</strong>, elimination<br />

of booster stations, and high hydraulic efficiency, which is su<strong>pe</strong>rior to centrifugal<br />

pumps.<br />

Plunger or diaphragm pumps do not handle large flow rates in excess of 100 m 3 /hr<br />

(4400 US gpm), but they are suitable for a wide range of applications at higher volumetric<br />

concentrations than centrifugal pumps. Positive displacement pumps can pump slurries<br />

with a weight concentration of 70%.<br />

9-1 SOLID PISTON PUMPS<br />

Positive displacement <strong>slurry</strong> pumps are used in a number of industries (Table 9-1). Solid<br />

piston pumps are reserved for the pumping of slurries of a low to medium abrasiveness<br />

(Miller Number


9.2 CHAPTER NINE<br />

TABLE 9-1 Applications of Positive Displacement Pumps<br />

Industry Application<br />

Mining Coal transportation (e.g., Novo Siberski pi<strong>pe</strong>line, Black Mesa Pi<strong>pe</strong>line)<br />

Flotation material<br />

Washery refuse<br />

Deep mine dewatering of water with solid particles<br />

Limestone, milk of lime<br />

Potash rock salt, phosphate, iron ore, nickel ore concentrate<br />

Bauxite, red mud, gold mud<br />

Sand, pyrite, REA gypsum<br />

Backfilling<br />

Underground drainage<br />

Filter press feed<br />

Autoclave feed<br />

Chemicals Salt <strong>slurry</strong><br />

Porcelain <strong>slurry</strong><br />

Pastes<br />

Detergent <strong>slurry</strong><br />

Combustion furnace feed<br />

Filter press feed<br />

Power generation Coal and coal <strong>slurry</strong><br />

Flue<br />

Pressurized fluidized bed combustion<br />

Wet ash removal<br />

Ship loading<br />

Long-distance pi<strong>pe</strong>lines<br />

Construction Bentonite, clay mash, cement<br />

Porcelain Clay <strong>slurry</strong><br />

Filter press feed<br />

Sugar Carbonation <strong>slurry</strong><br />

Sugar beet washing<br />

Information provided by courtesy of Wirth-Maschinen and Bohrgerate, Germany.<br />

TABLE 9-2 Comparison between Duplex and Triplex Pumps<br />

Duplex Single Acting Duplex Double Acting Triplex Single Acting<br />

2 cylinder liners 2 cylinder liners 3 cylinder liners<br />

2 piston gaskets 4 piston gaskets 3 piston gaskets<br />

2 cylinders 2 cylinders 3 cylinders<br />

2 piston rods with packing<br />

4 valves 8 valves 6 valves<br />

Slurry does not come in Slurry does come in contact Slurry does not come in<br />

contact with packing with packing contact with packing


POSITIVE DISPLACEMENT PUMPS<br />

FIGURE 9-1 Concept of the double-acting duplex piston pump. (Courtesy of Wirth<br />

Pumps.)<br />

the piston, these pumps o<strong>pe</strong>rate at lower s<strong>pe</strong>eds than single-acting duplex and triplex<br />

pumps (Figure 9-2). The single-acting duplex pump seems to have disap<strong>pe</strong>ared from the<br />

world of manufacturing.<br />

For pro<strong>pe</strong>r balancing, the pistons of duplex pumps are 180° out of phase (Figure 9-3),<br />

but for triplex pumps they are 120° out of phase with each other (Figure 9-4).<br />

Triplex pumps have a lower degree of oscillation than duplex units. The degree of<br />

FIGURE 9-2 Concept of the triplex piston pump. (Courtesy of Wirth Pumps.)<br />

9.3


9.4 CHAPTER NINE<br />

FIGURE 9-3 Concept of gear mechanism for duplex piston pumps. (Courtesy of Wirth<br />

Pumps.)<br />

FIGURE 9-4 Concept of gear mechanism for triplex piston pumps. (Courtesy of Wirth<br />

Pumps.)


POSITIVE DISPLACEMENT PUMPS<br />

FIGURE 9-5 Pulsation diagram for triplex pump. (Courtesy of Wirth Pumps.)<br />

variation of flow in the former is 23% (Figure 9-5) compared to 46% in the latter (Wallrafen,<br />

1983; Figure 9-6).<br />

Duplex and triplex <strong>slurry</strong> pumps are manufactured to a power frame of approximately<br />

1500 kW (2000 bhp). The Black Mesa Pi<strong>pe</strong>line featured 13 duplex pumps, each with a<br />

driving power of 1250 kW (1675 bhp) to transport 4.8 million tons of coal over a distance<br />

of 440 km (275 miles) (Wallrafen, 1983). Some of these pumps were manufactured by<br />

Wilson-Snyder in the United States.<br />

Piston <strong>slurry</strong> pumps are used extensively as mud transfer pumps. Gardner-Denver in<br />

the United States offers a range of duplex pumps in the power range of 12–76 kW<br />

(16–102 hp).<br />

FIGURE 9-6 Pulsation diagram for duplex pump. (Courtesy of Wirth Pumps.)<br />

9.5


9.6 CHAPTER NINE<br />

FIGURE 9-7 Triplex pump TPK 7� × 12�/1600. Driving power 1200 kW (1600 hp). (Courtesy<br />

of Wirth Pumps.)<br />

9-2 PLUNGER PUMPS<br />

For a long time, plunger pumps were not considered to be suitable for <strong>slurry</strong> transportation,<br />

but in the late 1970s, manufacturers develo<strong>pe</strong>d a suitable flushing system to minimize<br />

wear of the plunger.<br />

Plunger pumps are single acting. They use plungers instead of pistons (Figure 9-8)<br />

with valves. They o<strong>pe</strong>rate at 80–120 cycles <strong>pe</strong>r minute.<br />

Plunger pumps are prone to wear. They are less ex<strong>pe</strong>nsive to purchase than diaphragm<br />

pumps but have a higher maintenance cost. These pumps use three ty<strong>pe</strong>s of valves:<br />

1. Free-floating valves<br />

2. Spring-loaded spherical (Rollo) valves<br />

3. Spring-loaded elastomer-seal (mud) valves<br />

These valves are shown in Figure 9-9. It is important to minimize packing wear with<br />

piston pumps. Smith (1985) proposed four methods:<br />

1. Use of conventional packing of plungers at low s<strong>pe</strong>ed with slurries of low abrasiveness<br />

2. Provision of a clean, <strong>slurry</strong>-free environment for the packing rubbing surface (by synchronized<br />

or continuous injection of water or cleaning fluid)<br />

3. Separation of the <strong>slurry</strong> from the pumping element<br />

4. Total isolation of the <strong>slurry</strong> from the packing (by providing a separate diaphragm<br />

chamber)<br />

The SAMARCO pi<strong>pe</strong>line in Brazil used 14 plunger pumps with a driver power of 920<br />

kW each to deliver 12 million metric tons of iron oxide ore concentrate over a distance of


POSITIVE DISPLACEMENT PUMPS<br />

FIGURE 9-8 Schematic representation of a plunger pump.<br />

400 km (250 mi) (Wallrafen, 1983). These triplex plunger pumps were manufactured by<br />

Wilson-Snyder.<br />

The Wilson-Snyder line of plunger <strong>slurry</strong> pumps features 21 different sizes from<br />

45–1250 kW (60–1700 hp). They have also been used in Georgia, U.S.A. on a kaolin<br />

pi<strong>pe</strong>line. Kaolin is not very abrasive. These pumps have also been used for mine dewatering<br />

from a depth of 1036 m (3400 ft). The water contained solids.<br />

FIGURE 9-9 Categories of valves for <strong>slurry</strong> pumps. (a) Free floating; (b) spring-loaded<br />

spherical (Rollo); (c) spring-loaded elastomer seal. (From Smith, 1985. Reprinted by <strong>pe</strong>rmission<br />

of McGraw-Hill.)<br />

9.7


9.8 CHAPTER NINE<br />

Some of the triplex plunger pumps are rated at 41.4 MPa (6000 psi) (Wilson-Snyder,<br />

1977), such as the model 85-25. The volume capacity is rated from 363–2941 L/min<br />

(96–777 US gpm).<br />

9-3 PISTON DIAPHRAGM PUMPS<br />

To handle abrasive slurries that piston pumps would find difficult, manufacturers such as<br />

Geho Pumps (Netherlands), Wirth (Germany), and Gorman-Rupp (United States) have<br />

develo<strong>pe</strong>d pumps to use a diaphragm or a sort of flexible piston that comes in contact<br />

with the <strong>slurry</strong> or sludge. Feluwa of Germany added a hose so that there is an isolating<br />

bath of oil between the hose and the diaphragm. The pumps from Geho, Wirth, and<br />

Feluwa feature a crankshaft mechanism to move the diaphragm but the Gorman-Rupp<br />

pump uses an air cylinder to actuate the diaphragm.<br />

Diaphragm piston pumps use a sort of oil chamber between a reciprocating piston (o<strong>pe</strong>rated<br />

by a crankshaft) and the diaphragm (Figure 9-10). Provided that no puncture occurs<br />

in the diaphragm, <strong>slurry</strong> does not come in contact with the piston. A s<strong>pe</strong>cial control<br />

system is installed to detect diaphragm rupture.<br />

Wallrafen (1983) reported that Wirth manufactured its first piston diaphragm pump in<br />

1969 to pump sand slurries. The first unit lasted 1000 hrs without having to replace worn<br />

parts. By the early 1980s, wear life of 6000 hrs was achieved with rubber materials and<br />

FIGURE 9-10 Concept of double-acting piston duplex diaphragm pump. (Courtesy of Wirth<br />

Pumps.)


POSITIVE DISPLACEMENT PUMPS<br />

pro<strong>pe</strong>r design of the diaphragm. Diaphragm piston pumps are designed as duplex doubleacting<br />

or as triplex single-acting pumps in a similar concept rather to solid piston pumps<br />

(Figures 9-10 and 9-11).<br />

Piston diaphragm pumps (Figure 9-12) are more ex<strong>pe</strong>nsive than plunger pumps. For<br />

autoclave feed pumps, Geho develo<strong>pe</strong>d a s<strong>pe</strong>cial design to handle slurries as hot as<br />

200°C (392°F) at a high flow rate. Solids concentrations can be as high as75% and<br />

pumps can o<strong>pe</strong>rate at <strong>slurry</strong> tem<strong>pe</strong>ratures up to 200°C. Typical uses in the mining industry<br />

include:<br />

� Long-distance <strong>slurry</strong> (mineral concentrate) pi<strong>pe</strong>lines (up to 300 km long)<br />

� Clean and efficient tailings disposal<br />

� High-pressure bauxite digester feed<br />

� Autoclave and reactor feed<br />

� Mine backfilling<br />

� Mine dewatering (single stage)<br />

� Hydraulic ore hoisting<br />

With more than 400 piston diaphragm pumps installed on some of the world’s most demanding<br />

long-distance pi<strong>pe</strong>line applications, Geho Pumps has taken the opportunity,<br />

through a significant research and development effort, to constantly improve piston diaphragm<br />

pump design. This has resulted in numerous proprietary design improvements<br />

and technical innovations relevant to severe <strong>slurry</strong> pumping.<br />

FIGURE 9-11 Concept of single-acting piston triplex diaphragm pump. (Courtesy of Wirth<br />

Pumps.)<br />

9.9


9.10 CHAPTER NINE<br />

FIGURE 9-12 Geho pump at Freeport. (Courtesy of Geho.)<br />

Geho Piston diaphragm are used for tailings disposal and pumping at very high concentration<br />

so that:<br />

� Amount of free water is virtually eliminated, allowing <strong>slurry</strong> to be stacked or distributed<br />

in layers<br />

� Dry stacking requires less storage space<br />

� Prevents contamination of the environment by leakage<br />

� Rain and wind do not affect the solidified tailings<br />

� Mechanical stability of tailings allows a high stack with rehabilitation possibilities after<br />

use<br />

Typical tailings applications of Geho pumps include:<br />

� Bayswater Power Station—fly ash disposal (Australia)<br />

� Nabalco, Gove Refinery—red mud disposal (Australia)<br />

� Ledvice Power Station—fly ash disposal (Czech Republic)<br />

� Pingguo Aluminium Company—red mud disposal (Peoples Republic of China)<br />

� Kha<strong>pe</strong>rkheda Ash Handling Plant—fly ash disposal (India)<br />

� National Aluminum Company—red mud disposal (India)<br />

TABLE 9-3 Examples of Installation of Piston Diaphragm Pumps on Slurry Pi<strong>pe</strong>line<br />

and Tailings Applications<br />

Location Manufacturer Installation application flow rate stated for each pump<br />

Alsen Zementwerke Geho 3 piston diaphragm pumps to pump limestone <strong>slurry</strong> over<br />

10 km (6.3 mi)<br />

Antamina, Peru Wirth 4 piston pumps, 100 m3 /hr (440 gpm), 25.2 MPa (3650<br />

psi), cop<strong>pe</strong>r concentrate<br />

Ashanti Goldfields Ghana Wirth 1 pump, 215 m3 /hr (947 gpm), 6 MPa (870 psi) backfill<br />

slime (gold tailings)


TABLE 9-3 (continued)<br />

POSITIVE DISPLACEMENT PUMPS<br />

9.11<br />

Location Manufacturer Installation application flow rate stated for each pump<br />

Bajo Alumbrera, Argentina Geho 6 piston diaphragm pumps in 3 booster stations to<br />

transport cop<strong>pe</strong>r concentrate over a distance of 320 km<br />

(225 mi) in a 150 mm (6 in) line, 91 m 3 /h at 217 bar<br />

Cameco, Canada Wirth 2 pumps, 80 m 3 /hr (352 gpm), 12.5 MPa (1810 psi),<br />

uranium ore<br />

Cia Minera Disputada de Wirth 3 pumps 115 m 3 /hr (510 gpm), 2.5 MPa (360 psi) cop<strong>pe</strong>r<br />

Las Condes, Chile tailings<br />

Codelco, Chile Wirth 3 pumps, 140 m 3 /hr (616 gpm), 3.5 MPa (507 psi), c<br />

op<strong>pe</strong>r tailings<br />

Course Nickel , Australia Wirth 2 pumps, 193 m 3 /hr (850 gpm), 5 MPa (725 psi), lateritic<br />

nickel ore<br />

Doña Ines Collahuasi, Chile Geho 2 piston diaphragm pumps for 203 km of cop<strong>pe</strong>r<br />

concentrate transport, 117 m 3 /h at 217 bar<br />

ECPSA, Cuba Wirth 10 pumps, 200 m 3 /hr (800gpm), 6.4 MPa (920 psi),<br />

iron–nickel <strong>slurry</strong><br />

Empresa Minera Yauliyacu, Wirth 2 pumps, 90 m 3 /hr (400 gpm), 14.4 MPa (2080 psi),<br />

Peru tailings<br />

Eskay Creek, Canada Wirth 1 pump, 25 m 3 /hr (110 gpm), 11.7 MPa (1700 psi) for a<br />

6.5 km (4 mi) pi<strong>pe</strong>line<br />

Freeport, Indonesia Geho 2 piston diaphragm pumps to transport cop<strong>pe</strong>r<br />

concentrate over 120 km (75mi), 159 m 3 /h at 40 bar<br />

Goldmine, South Africa Wirth 1 pump, 12 m 3 /hr (66 gpm), 12 MPa (1714 psi),<br />

backfilling<br />

ISCOR, Hillendake Mine, Wirth 2 pumps 450 m 3 /hr (1980 gpm), 7.4 MPa (1075 psi),<br />

South Africa heavy mineral tailings<br />

Jian Shan Geho 100 kms iron ore concentrate transport (PR of China), 2<br />

piston diaphragm pumps, 216 m 3 /h at 153 bar<br />

Nabalco, Australia Geho 3 piston diaphragm pumps for a highly concentrated red<br />

mud <strong>slurry</strong> to a disposal area, 200 m 3 /h at 160 bar<br />

Norilsk Nickel Combinat Geho 9 piston diaphragm pumps, 55 km of multimetallic ore<br />

transport (North Siberia, Russia), 400 m 3 /h at 80 bar<br />

Pasminco, Australia Wirth 3 pumps, 161 m 3 /hr (710 gpm), 12.5 MPa (1810 psi) zinc<br />

and lead concentrate<br />

Los Pelambres, Chile Geho 2 piston diaphragm pumps for 120 km pi<strong>pe</strong>line<br />

transportation of cop<strong>pe</strong>r concentrate <strong>slurry</strong>, 165 m 3 /h<br />

at 150 bar<br />

Sicartsa, Mexico Geho 1 piston diaphram pump for iron ore concentrate <strong>slurry</strong>,<br />

380 m 3 /h at 110 bar<br />

J. R. Simplot, USA Geho 4 piston diaphragm pumps for 100 km transportation of<br />

phosphate <strong>slurry</strong>, 97 m 3 /h at 228 bar<br />

Batu Hijau, Indonesia Geho 2 piston diaphragm pumps for 120 km cop<strong>pe</strong>r<br />

concentrate <strong>slurry</strong> transportation, 123 m 3 /h at 228 bar<br />

Cockburn Cement, Australia Geho 3 piston diaphragm pumps for shell and <strong>slurry</strong> transport,<br />

206 m 3 /h at 65 bar<br />

New Zealand Steel Geho 4 piston diaphragm pumps for ironsand concentrate<br />

transportation, 194 m 3 /h at 100 bar<br />

Rio Capim, Brazil Geho 2 piston diaphragm pumps for kaolin <strong>slurry</strong><br />

transportation, 293 m 3 /h at 57 bar<br />

Minera Escondid, Chile Geho 1 piston diaphragm pum for cop<strong>pe</strong>r concentrate <strong>slurry</strong><br />

transportation, 295 m 3 /h at 69 bar<br />

OEMK, Ukraine Geho 4 piston diaphragm pumps for iron ore <strong>slurry</strong><br />

transportation, 540 m 3 /h at 74 bar


9.12 CHAPTER NINE<br />

Geho piston diaphragm pumps are suitable for feeding autoclaves with ore <strong>slurry</strong> in different<br />

mineral processes, such as in the aluminum, gold, and nickel industries. A s<strong>pe</strong>cial<br />

design of the Geho piston diaphragm pump is develo<strong>pe</strong>d for “hot” slurries. As the diaphragm<br />

cannot be exposed to high tem<strong>pe</strong>ratures, Geho Pumps develo<strong>pe</strong>d a dropleg concept,<br />

which allows transfer of hot <strong>slurry</strong> without having to cool the <strong>slurry</strong> down. Thus, the<br />

proposed pump design excels in low maintenance cost, low energy cost, and high reliability.<br />

Geho Pumps designed the dropleg concept in the early 1980s for feeding gold ore slurries<br />

to autoclaves for pressure oxidation at installations in Nevada, U.S.A. Recently, Geho<br />

Pumps develo<strong>pe</strong>d an improved dropleg configuration for 200°C slurries. These developments<br />

include, for example, a horizontal dropleg layout, a patented separator, improved<br />

dropleg efficiency, patented slide mounting of the pump to com<strong>pe</strong>nsate for thermal expansion,<br />

etc. Typical examples of high-tem<strong>pe</strong>rature autoclave feeding using Geho pumps<br />

include:<br />

� Bulong Nickel project—200°C laterite nickel <strong>slurry</strong> (Australia)<br />

� Murrin Murrin—200°C laterite nickel <strong>slurry</strong> (Australia)<br />

� American Barrick Phases I, II, and III—gold <strong>slurry</strong> (United States)<br />

� Twin Creeks Phases I and II—gold <strong>slurry</strong> (United States)<br />

Diaphragm piston pumps use two ty<strong>pe</strong>s of valves:<br />

1. Ball valves<br />

2. Conical valves<br />

The ball valves are a kind of check valve on suction and discharge and move according to<br />

outlet ball valve<br />

air exhaust<br />

Air<br />

<strong>slurry</strong><br />

air piston<br />

air inlet<br />

diaphragm<br />

FIGURE 9-13 Concept of air o<strong>pe</strong>rated <strong>slurry</strong> diaphragm pump.<br />

inlet ball valve


POSITIVE DISPLACEMENT PUMPS<br />

fluid forces to close and o<strong>pe</strong>n the inlet and the outlet. Their own weight plays a role in<br />

o<strong>pe</strong>ning and closing. They are more suitable for the lower end of pressures. When the<br />

pressures are very high, the sealing of the ball valve cannot support the force. Conical<br />

spring-actuated valves are then installed; the force of the spring helps to close the valves.<br />

Conical valves allow o<strong>pe</strong>ration at higher pressures due to metal-to-metal support.<br />

Air driven diaphragm pump use a piston connected to the diaphragm. In addition, air<br />

can be forced against the dry side of the diaphragm to help move the diaphragm. Air<br />

leaves at the top piston assembly. The pump uses ball valves for the inlet and exit of the<br />

<strong>slurry</strong> (Figure 9-13). The air-o<strong>pe</strong>rated diaphragm pumps serve a niche of the market and<br />

have a range of discharge from 15–100 mm (5/8–4 in) and flows up to 9 L/s (150 US<br />

gpm). The maximum head from these pumps is of the order of 15 m (50 ft).<br />

9-4 ACCESSORIES FOR PISTON AND<br />

PLUNGER PUMPS<br />

9.13<br />

The pulsations of a positive displacement pump are transmitted to the <strong>slurry</strong>. To prevent<br />

their propagation to the pi<strong>pe</strong>line and its support, it is essential to install hydraulic dam<strong>pe</strong>ners<br />

(Figure 9-14).<br />

A hydraulic dam<strong>pe</strong>ner is essentially a chamber with a diaphragm. On one side there is<br />

<strong>slurry</strong> and on the other there is a gas such as nitrogen under compression to absorb the oscillations.<br />

Dam<strong>pe</strong>ners are installed on the discharge of the pump, and in some cases on the<br />

suction too.<br />

The manufacturers of piston and diaphragm pumps provide a package of s<strong>pe</strong>cial tools<br />

to install replacement parts.<br />

FIGURE 9-14 Hydraulic dam<strong>pe</strong>ner for diaphragm pumps. (Courtesy of Wirth Pumps.)


9.14 CHAPTER NINE<br />

FIGURE 9-15 Peristaltic pump. (Courtesy of Gorman Pump Industries.)<br />

Air o<strong>pe</strong>rated diaphragm pumps are noisy and need a silencer on the exhaust of the air.<br />

9-5 PERISTALTIC PUMPS<br />

A <strong>pe</strong>ristaltic pump is essentially a hose that is pressed by a cam, an eccentric mechanism,<br />

or three rollers on arms (Figure 9-15). The pressure is then transmitted to the fluid. They<br />

are available in a range of flow from microliters/min up to 33 L/min (8.8 gpm) and pressures<br />

up to 420 kPa (60 psi). They are popular for medical applications as they do not<br />

cause damage to blood cells. Peristaltic pumps are used to transport highly concentrated<br />

<strong>slurry</strong> at a small flow rate for a s<strong>pe</strong>cific range of applications such as clay, gold, and platinum<br />

slurries, and filter press feed. They are self-priming and develop a high vacuum, up<br />

to 635 mm (25 in) on the mercury scale.<br />

9-6 ROTARY LOBE SLURRY PUMPS<br />

Rotary lobe pumps are a s<strong>pe</strong>cial form of positive displacement pump. They feature two<br />

lobes (Figure 9-16) that rotate against each other like intermeshing gears. They are avail-


POSITIVE DISPLACEMENT PUMPS<br />

able for flow rates in the range of 0–170 L/s (0270 US gpm) and for discharge pressures<br />

up to 1.2 MPa (175 psi). They are self-priming up to a negative suction of 8 m (24 in) and<br />

can handle viscous and abrasive slurries. They are capable of dry running up to 30 min.<br />

The lobes for <strong>slurry</strong> handling are made from abrasion-resistant alloys or a steel core with<br />

molded rubber surfaces. The casing is hardened.<br />

Rotary lobe <strong>slurry</strong> pumps are used with certain soft slurries with mild abrasion characteristics<br />

such as wastewater and sewage disposal, flotation slimes, digested scum, lime<br />

<strong>slurry</strong> in waste treatment plants. They are also used in food processing to move potato and<br />

starch pulp, mash, food paste, tomato paste, food wastes, and dairy waste and whey in milk<br />

processing. They are used in the pa<strong>pe</strong>r industry to pump lime <strong>slurry</strong> and adhesives. They<br />

are used in the plastic recycling industry to pump plastic and Styrofoam cups. They are also<br />

used in the construction industry for pumping bentonite slurries, clay slurries, and mud.<br />

9-7 THE LOCKHOPPER PUMP<br />

9.15<br />

FIGURE 9-16 Rotary positive displacement pump. (Courtesy of Gorman Pump Industries.)<br />

The lockhop<strong>pe</strong>r is not exactly a pump, but rather a system to pump <strong>slurry</strong>, including fairly<br />

coarse material at extremely high pressure. It consists of two chambers that alternate in


9.16 CHAPTER NINE<br />

water tank<br />

Multi-stage<br />

water pump<br />

injecting water and <strong>slurry</strong> into the pi<strong>pe</strong>line. Injection is provided by water pressure. The<br />

water is separated from the <strong>slurry</strong> by a diaphragm in the form of a free-rolling rubber<br />

spherical “piston,” a sort of double-acting piston with water on one side and <strong>slurry</strong> on the<br />

other. On one side of the piston, <strong>slurry</strong> enters from feeding hop<strong>pe</strong>rs and is charged to the<br />

pi<strong>pe</strong>line. On the other side of the piston, water enters—pum<strong>pe</strong>d by high-pressure, multistage<br />

water pumps—which pro<strong>pe</strong>ls the piston before being returned to water tanks. Figure<br />

9-17 illustrates the lockhop<strong>pe</strong>r system. The lockhop<strong>pe</strong>r can be adapted to pump 2� coarse<br />

coal, bauxite lumps, and other materials whose size would be beyond the range of diaphragm<br />

and plunger pumps.<br />

Although the lock hop<strong>pe</strong>r can be designed for very high pressures, the limitation is on<br />

the practical size of pi<strong>pe</strong>line pressure rating that can be selected. This is often of the order<br />

of 21 MPa (3000 psi).<br />

9-8 CONCLUSION<br />

<strong>slurry</strong> hop<strong>pe</strong>r<br />

Valve<br />

Water Slurry Valve<br />

Valve<br />

Valve<br />

FIGURE 9-17 Concept of the lockhop<strong>pe</strong>r system.<br />

Slurry Pi<strong>pe</strong>line<br />

free rolling piston<br />

Positive displacement pumps play a major role in the power industry and the pumping of<br />

highly concentrated slurries and food and chemical pastes at high efficiency. Considerable<br />

development in the last quarter of the 20th century by manufacturers of these pumps,<br />

particularly diaphragm pumps, has led to a wide range of applications from long-distance<br />

pi<strong>pe</strong>line pumping stations, to mine dewatering, to pumping concrete with high degree of<br />

reliability. Research continues in the field of high-tem<strong>pe</strong>rature applications such as autoclave<br />

feeds.


9-9 REFERENCES<br />

POSITIVE DISPLACEMENT PUMPS<br />

9.17<br />

Wallrafen, G. 1983. Piston Pumps for the Hydraulic Transport of Solids. Bulk Solids Handling, 3, 1.<br />

Wallrafen, G. 1985. Backfilling with Viscous Slurry Pumps. Bulk Solids Handling 5, no. 4: Wilson-<br />

Snyder. 1977. Slurry Pumps. Texas: Wilson Snyder. Publication ADWS 28-77 (3M).<br />

Smith W.1985. Construction of Solids-Handling Displacement Pumps. Chapter 9-17-3 in The Pump<br />

Handbook, Karassik I. J., W. C. Krutzch, W. H. Fraser, and J. P. Messina (Eds.). New York:<br />

McGraw-Hill.


CHAPTER 5<br />

HOMOGENEOUS FLOWS<br />

OF NONSETTLING<br />

SLURRIES<br />

5-0 INTRODUCTION<br />

The rheology of non-Newtonian flows and homogeneous flows was examined in detail in<br />

Chapter 3. With modern methods of grinding, the size of particles can be reduced to values<br />

smaller than 70 �m (0.0028 in). As shown in Table 3-8, a wide range of metal concentrates<br />

and tailings are pum<strong>pe</strong>d at a sufficiently high concentration with a small enough<br />

particle size for the mixture to behave as a Bingham plastic. There are other clays and<br />

slurries that may behave as pseudoplastics. In certain circuits of oil sands processing<br />

plants with tar at the level of flotation, the <strong>slurry</strong> may behave as a thixotropic mixture.<br />

With non-Newtonian flows, it is important to take into account the rheology, yield<br />

stress, power law exponent, coefficient, and even the time response. Different models<br />

have evolved over the years for Bingham and pseudoplastic slurries. Some of these models<br />

put more emphasis on the laminar flow regime, in which roughness effects are negligible.<br />

Some other models extend to the transition and turbulent regimes. The effect of pi<strong>pe</strong><br />

roughness on friction loss factors in non-Newtonian flows remains a topic worth investigating<br />

and researching.<br />

For thixotropic slurries, methods are used to predict start-up pressure after a shutdown<br />

of the pi<strong>pe</strong>line, and the time required to clean the conduit of gelled material before resuming<br />

pumping.<br />

The equations for friction factors of non-Newtonian fluids are fairly complex and require<br />

iteration and data on the rheology. Throughout the years, different authors have develo<strong>pe</strong>d<br />

equations for “modified” Reynolds numbers, Hedstrom numbers, etc. In this<br />

chapter, equations develo<strong>pe</strong>d by different authors will be reviewed. Through worked examples,<br />

the reader will be shown methods of calculating the friction factor. It is the purpose<br />

of this chapter to focus on the engineering side of the problem. The reader is strongly<br />

advised to read through Chapter 3, to gain the fundamentals for this chapter.<br />

For the practical engineer, who is more concerned with the actual design of a pi<strong>pe</strong>line<br />

or a pumping system, the numerous and different definitions of the so-called “modified<br />

Reynolds number” can be very confusing. Every few years, an author develops a new definition<br />

of the “modified Reynolds number” and claims to have found a relationship with<br />

the friction factor. The cautious approach for an engineer is to assume that such a universal<br />

relationship is illusive, and for every ty<strong>pe</strong> of <strong>slurry</strong> there may be a model to use.<br />

5.1


5.2 CHAPTER FIVE<br />

Sometimes the use of two different methods can yield differences of 25–35% in the estimation<br />

of the friction factor.<br />

One important difference with the slurries described in Chapter 4 is that most non-<br />

Newtonian slurries do not exhibit the stratification of solids, which is usual with coarse<br />

particles.<br />

5-1 FRICTION LOSSES FOR<br />

BINGHAM PLASTICS<br />

Bingham plastics were defined in Chapter 3. They are characterized by a yield stress that<br />

must be overcome to start the flow. Examples of Bingham plastics are listed in Table 3-9<br />

5-1-1 Start-up Pressure<br />

The start-up pressure for pumping a Bingham plastic is expressed in terms of the true<br />

yield stress:<br />

4�0L Pst = � (5-1)<br />

Di<br />

The start-up pressure <strong>pe</strong>r unit length is obtained by dividing Equation 5-1 by the<br />

length of the pi<strong>pe</strong>line:<br />

P st<br />

� L<br />

=<br />

Examples for the starting pressure <strong>pe</strong>r unit length for slurries in 3�, 6�, 12�, and 18�<br />

pi<strong>pe</strong>s are presented in Table 5-1.<br />

The Reynolds Number for a Bingham <strong>slurry</strong> is expressed as:<br />

The coefficient of rigidity � was defined in Equation 3-29 as<br />

ReB = � (5-2)<br />

�<br />

� 0<br />

4� 0<br />

� Di<br />

D iV� m<br />

� = � + �� (3-29)<br />

(d�/dt)<br />

For Bingham slurries a nondimensional coefficient is defined as the plasticity number:<br />

�0Di PL = � (5-3)<br />

�V<br />

The Hedstrom number is the product of the plasticity number and the Reynolds number<br />

and is calculated as<br />

2 D i �m�0 �2 �<br />

He = (5-4)<br />

Table 5-2 shows examples of Bingham <strong>slurry</strong> mixtures and the magnitude of the Hedstrom<br />

number for flows in rubber-lined 6� (150 mm NB), 12� (300 mm NB) and 18� (450


5.3<br />

TABLE 5-1 Starting Pressure <strong>pe</strong>r Unit Length for Certain Slurries in Pa/m<br />

Starting pressure <strong>pe</strong>r unit length (Pa/m)*<br />

3� Pi<strong>pe</strong> 6� Pi<strong>pe</strong><br />

Coefficient Sch 40, Sch 40, 12� Pi<strong>pe</strong> S, 18� Pi<strong>pe</strong> S,<br />

Yield of rigidity, rubber lined, rubber lined, rubber lined, rubber lined,<br />

Particle size, Density, stress, � mPa · s ID = 2.560� ID = 5.557� ID = 11.500� ID = 16.500�<br />

Slurry d 50 kg/m 3 Pa (cP) (65 mm) (141.2 mm) (292 mm) (419 mm)<br />

54.3% Aqueous sus<strong>pe</strong>nsion of 92% under 74 �m 1520 3.8 6.86 234 108 52 37<br />

cement rock<br />

Flocculated aqueous China clay 80% under 1 �m 1280 59 13.1 3631 1671 808 567<br />

sus<strong>pe</strong>nsion No. 1<br />

Flocculated aqueous China clay 80% under 1 �m 1207 25 6.7 1538 708 342 240<br />

sus<strong>pe</strong>nsion No. 4<br />

Flocculated aqueous China clay 80% under 1 �m 1149 7.8 4.0 480 221 107 75<br />

sus<strong>pe</strong>nsion No. 6<br />

Aqueous clay sus<strong>pe</strong>nsion I 1520 34.5 44.7 2123 977 473 332<br />

Aqueous clay sus<strong>pe</strong>nsion III 1440 20 32.8 1231 567 274 192<br />

Aqueous clay sus<strong>pe</strong>nsion V 1360 6.65 19.4 409 188 91 64<br />

21.4% Bauxite


5.4<br />

TABLE 5-2 Examples of Hedstrom Numbers for Bingham Slurries in 3�, 6�, 12�, and 18� Rubber-Lined Pi<strong>pe</strong>s<br />

Hedstrom number*<br />

3� Pi<strong>pe</strong> 6� Pi<strong>pe</strong><br />

Coefficient Sch 40, Sch 40, 12� Pi<strong>pe</strong> S, 18� Pi<strong>pe</strong> S,<br />

Yield of rigidity, rubber lined, rubber lined, rubber lined, rubber lined,<br />

Particle size, Density, stress, � mPa · s ID = 2.560� ID = 5.557� ID = 11.500� ID = 16.500�<br />

Slurry d 50 kg/m 3 Pa (cP) (65 mm) (141.2 mm) (292 mm) (419 mm)<br />

54.3% Aqueous sus<strong>pe</strong>nsion of 92% under 74 �m 1520 3.8 6.86 518,568 2,445,272 1.046 × 10 7 2.124 × 10 7<br />

cement rock<br />

Flocculated aqueous China clay 80% under 1 �m 1280 59 13.1 1,859,285 8,773,821 3.752 × 10 7 7.616 × 10 7<br />

sus<strong>pe</strong>nsion No. 1<br />

Flocculated aqueous China clay 80% under 1 �m 1207 25 6.7 2,840,039 1.472 × 10 7 6.078 × 10 7 1.234 × 10 8<br />

sus<strong>pe</strong>nsion No. 4<br />

Flocculated aqueous China clay 80% under 1 �m 1149 7.8 4.0 2,366,581 1.117 × 10 7 4.776 × 10 7 9.694 × 10 7<br />

sus<strong>pe</strong>nsion No. 6<br />

Aqueous clay sus<strong>pe</strong>nsion I 1520 34.5 44.7 110,885 523,259 2,237,759 4,541,865<br />

Aqueous clay sus<strong>pe</strong>nsion III 1440 20 32.8 113,102 533,721 2,282,499 4,632,671<br />

Aqueous clay sus<strong>pe</strong>nsion V 1360 6.65 19.4 101,528 479,100 2,048,910 4,158,568<br />

21.4% Bauxite


mm) pi<strong>pe</strong>s. Laminar flows in large pi<strong>pe</strong>s are considered for certain slurries at relatively<br />

high weight concentration (� 60%), or certain high-energy mixtures (e.g., crude oil with<br />

fine and ultrafine coal).<br />

Example 5-1<br />

A <strong>slurry</strong> consists of a clay and water mixture. It is tested and classified as a Bingham mixture<br />

with a yield stress of 17 Pa. The pi<strong>pe</strong> inner diameter is 63 mm. The pi<strong>pe</strong> length is<br />

6500 m. Determine the start-up pressure, ignoring any static head.<br />

Solution in SI Units<br />

Using Equation 5-1:<br />

Solution in US units<br />

HOMOGENEOUS FLOWS OF NONSETTLING SLURRIES<br />

P st = 4� 0L/D i<br />

P st = 4 × 17 × 6500/0.063 = 7,015,873 (1018 psi)<br />

P st = 4� 0L/D i<br />

� 0 = 17 Pa/6895 = 2.465 × 10 –3 psi<br />

L = 6500 m/0.0254 = 255,905 in<br />

Di = 63/25.4 = 2.48 in<br />

4 × 2.465 × 10<br />

Start-up pressure Pst = = 1017.6 psi<br />

–3 × 255,905<br />

���<br />

2.48<br />

5-1-2 Friction Factor in Laminar Regime<br />

Buckingham (1921) was the first to develop an equation for a fully develo<strong>pe</strong>d laminar<br />

flow. This equation has since been modified by Hedstrom (1952) and others to express<br />

the friction factor as a function of the Hedstrom and Reynolds numbers:<br />

= – + (5-5)<br />

or<br />

fNL = �1 + – He<br />

� (5-6)<br />

This would occur below the critical Reynolds number or in transition between laminar<br />

and turbulent flow. The last term between brackets in the equation is often considered<br />

second order.<br />

4<br />

He<br />

16 He<br />

� � �3 7<br />

ReB 6ReB 3 f NLReB 4<br />

1 fNL He<br />

� � � � 2 3 8<br />

ReB 16 6ReB 3 f NLRe B<br />

Example 5-2<br />

A Bingham <strong>slurry</strong> with a concentration of 50% by weight is tested in a plastic-lined pi<strong>pe</strong><br />

with an inner diameter of 2.5 in. The tests indicate a yield stress of 1.5 Pa, a <strong>slurry</strong> mixture<br />

s<strong>pe</strong>cific gravity of 1.54, and a coefficient of rigidity of 0.4 Pa · s. Assuming a flow<br />

s<strong>pe</strong>ed of 4 ft/s in a laminar regime, determine the friction factor by Buckingham’s equation.<br />

5.5


5.6 CHAPTER FIVE<br />

Solution in SI Units<br />

Pi<strong>pe</strong> ID = 2.5� or 63.5 mm<br />

S<strong>pe</strong>ed = 4 ft/s or 1.219 m/s<br />

Reynolds number = 0.0635 × 1.219 × 1540/0.4 = 298<br />

Hedstrom number = (0.0635) 2 × 1540 × 1.5/(0.4 2 ) = 916.8<br />

Using Equation 5-6 and ignoring the higher-order terms:<br />

f NL � [16/Re B][1 + He/(6Re B)] = 0.0812<br />

This a fanning factor, so the Darcy friction factor is 0.3248.<br />

Figure 5-1 presents values of the friction factor versus the Reynolds number for a wide<br />

range of Hedstrom numbers from 0 to 10 9 . The transition between laminar and turbulent<br />

flows is shown by the dotted curve of the critical Reynolds number. From the point of<br />

view of engineering, the most practical flows in pi<strong>pe</strong>s are in a range of Hedstrom numbers<br />

between 10 5 and 10 8 , as shown in Table 5-2. The transition to turbulent flow will be examined<br />

in more detail throughout this chapter. To appreciate the practical magnitude of<br />

the laminar friction factor, Table 5-3 presents cases at a s<strong>pe</strong>ed of 1 m/s (3.3 ft/sec) for rubber-lined<br />

pi<strong>pe</strong>s in sizes of 3� (80 mm N.B.), 6� (150 mm N.B), 12� (300 mm N.B.), and<br />

18� (450 mm N.B.). The Fanning friction factor based on the Buckingham equation is in<br />

the range of 0.001 to 0.15.<br />

FIGURE 5-1 Friction factor versus Reynolds number and Hedstrom number. (From Hill R.<br />

A, P. E. Snoek, and R. L. Ghandi, Hydraulic transport of solids, in The Pump Handbook, 2nd<br />

ed., I. J. Karassik et al. (Eds.), New York: McGraw-Hill. Reproduced by <strong>pe</strong>rmission.)


5.7<br />

TABLE 5-3 Friction Factor at a S<strong>pe</strong>ed of 1 m/s (3.3 ft/sec) for Bingham Mixtures in Rubber-Lined Pi<strong>pe</strong>s*<br />

Fanning friction factor f N at a s<strong>pe</strong>ed of 1 m/s (3.3 ft/sec)<br />

3� Pi<strong>pe</strong> 6� Pi<strong>pe</strong><br />

Coefficient Sch 40, Sch 40, 12� Pi<strong>pe</strong> S, 18� Pi<strong>pe</strong> S,<br />

Yield of rigidity, rubber lined, rubber lined, rubber lined, rubber lined,<br />

Particle size, Density, stress, � mPa · s ID = 2.560� ID = 5.557� ID = 11.500� ID = 16.500�<br />

Slurry d 50 kg/m 3 Pa (cP) (65 mm) (141.2 mm) (292 mm) (419 mm)<br />

54.3% Aqueous sus<strong>pe</strong>nsion of 92% under 74 �m 1520 3.8 6.86 0.00778 0.00718 0.00691 0.00684<br />

cement rock<br />

Flocculated aqueous China clay 80% under 1 �m 1280 59 13.1 0.12541 0.12407 0.12348 0.12331<br />

sus<strong>pe</strong>nsion No. 1<br />

Flocculated aqueous China clay 80% under 1 �m 1207 25 6.7 0.00566 0.05586 0.05554 0.05545<br />

sus<strong>pe</strong>nsion No. 4<br />

Flocculated aqueous China 80% under 1 �m 1149 7.8 4.0 0.0186 0.0185 0.0183 0.0182<br />

sus<strong>pe</strong>nsion No. 6<br />

Aqueous clay sus<strong>pe</strong>nsion I 1520 34.5 44.7 0.0677 0.0639 0.0621 0.0616<br />

Aqueous clay sus<strong>pe</strong>nsion III 1440 20 32.8 0.0426 0.0396 0.0382 0.0379<br />

Aqueous clay sus<strong>pe</strong>nsion V 1360 6.65 19.4 0.01655 0.0146 0.01382 0.01359<br />

21.4% Bauxite


5.8 CHAPTER FIVE<br />

5-1-3 Transition to Turbulent Flow Regime<br />

Hanks and Pratt (1967) analyzed extensive ex<strong>pe</strong>rimental data on critical Reynolds numbers<br />

and proposed a relationship between the Reynolds and Hedstrom Numbers at transition as<br />

ReBc = �1 – x He 4 1<br />

� � 4<br />

c + � x c� (5-7)<br />

xc 3 3<br />

where xc = �0/�wc = the ratio of the yield stress to the wall shear stress at the transition<br />

from laminar to turbulent flow.<br />

At the transition,<br />

He = 16,800 � (5-8)<br />

(1 – xc) 3<br />

Figure 5-2 plots the magnitude of critical Reynolds number versus the Hedstrom number<br />

for a number of Bingham slurries.<br />

Example 5-3<br />

Using Figure 5-2, determine the critical Reynolds number for a clay <strong>slurry</strong> at a Hedstrom<br />

number of 10,000.<br />

Solution<br />

From Figure 5-2 Rec � 3700.<br />

Wasp et al. (1977) defined the effective pi<strong>pe</strong>line viscosity for laminar flow as<br />

Critical Reynolds Number<br />

10 5<br />

10 4<br />

10<br />

3<br />

3 4<br />

10 10<br />

x c<br />

�w �e = � (5-9)<br />

8V/Di<br />

5<br />

10<br />

6<br />

10<br />

8<br />

10<br />

Hedstrom Number<br />

FIGURE 5-2 The critical Reynolds number versus the Hedstrom number for flow in pi<strong>pe</strong>s.<br />

(After Hanks, R. W., and D. R. Pratt. 1967.)


HOMOGENEOUS FLOWS OF NONSETTLING SLURRIES<br />

In Chapter 3, in the description of the capillary tube test, the Buckingham equation was<br />

derived. When ignoring the fourth-power term, the equation reduces to<br />

�w � 8V�/Di + 4/3 �0 (3-62)<br />

or<br />

�e � ���1 + 8V�/Di + 4/3 �0 DI�0 �� �� (5-10)<br />

8V/Di<br />

6V�<br />

In many laminar flow pi<strong>pe</strong>lines the term Di�0/(6V�) is much smaller than unity. In such<br />

cases Equation 5-10 is then simplified to<br />

�e � Di�0/(6V) (5-11)<br />

At transition, Wasp et al. (1977) point out that this equation can be approximated to<br />

�e � [Di�0/(6VTR)] using the effective pi<strong>pe</strong>line viscosity, the critical Reynolds number is expressed as<br />

ReBC = 6V 2 TR�/�0 At high shear rate for Bingham plastics, � e � � � � � (see Figure 3-10):<br />

ReBC�0 VTR = � �� 6�<br />

(5-12)<br />

The Wasp method is based on numerous assumptions, and usually terminates by assuming<br />

that the transition Reynolds number is in the range of 2000 to 3000 for numerous<br />

Bingham slurries. In some res<strong>pe</strong>cts, it is a useful tool for hand calculations. A more widely<br />

accepted method since the mid 1980s is to compute the transition velocity that was proposed<br />

by Wilson and Thomas, to be discussed in Section 5-4-3. It is then assumed that the<br />

transition from laminar to turbulent flow occurs when the Wilson–Thomas and the Buckingham<br />

equations intersect.<br />

5-1-4 Friction Factor in the Turbulent Flow Regime<br />

Hanks and Dadia (1971) develo<strong>pe</strong>d a semiempirical equation for the turbulent flow of<br />

Bingham slurries in closed conduits. These equations were modified by Darby (1981) and<br />

Darby et al. (1992) to give a friction factor for the turbulent regime as<br />

fNT = 10a b ReB (5-13)<br />

where<br />

a = –1.47[1 + 0.146 exp(–2.9 × 10 –5He)] b = –0.193<br />

The values of the parameters a and b are based on empirical data for closed conduits.<br />

Bingham slurries do not exhibit a sudden change from laminar to turbulent flow. Darby<br />

et al. (1992) reviewed the work of previous authors and proposed to combine the laminar<br />

and turbulent fanning friction factors into the following equation:<br />

fN = ( f m NL + f m NT) (1/m) (5-14)<br />

where<br />

m = 1.7 + 40,000/ReB. (5-15)<br />

5.9


5.10 CHAPTER FIVE<br />

Equations 5-12 and 5-13 do not account for pi<strong>pe</strong> roughness and are essentially for very<br />

smooth pi<strong>pe</strong>s such as glass and high-density polyethylene pi<strong>pe</strong>s. Studies have been published<br />

in the past on flow of Bingham plastics in the laminar regime, where roughness effects<br />

are neglected. Thomas and Wilson (1987) have even argued that the non-Newtonian<br />

fluids form a viscous sublayer in the boundary layer, that is usually thicker than with<br />

Newtonian flows. This viscous sublayer is considered to suppress the contribution of<br />

roughness and, in effect, Wilson and Thomas do support the assumptions made by Darby<br />

(1981). Many concentrates are pum<strong>pe</strong>d at high volumetric concentrations but also at a relatively<br />

moderate s<strong>pe</strong>ed of 2–2.5 m/s (6.6–8.2 ft/s). We will, however, discuss the effects<br />

of roughness in Section 5-7.<br />

Example 5-4<br />

In Figure 3-9, the relationship between the Bingham plastic apparent viscosity and the<br />

shear rate was presented as<br />

� = �(d�/d�) + �� at high shear rate � � ��. Considering an aqueous clay sus<strong>pe</strong>nsion III (Table 3-9) with a mixture density of 1440<br />

kg/m3 (SG = 1.44), a yield stress of 20 Pa, and a coefficient of rigidity of 32.8 mPa · s as<br />

reported by Caldwell and Babbitt (see references of Chapter 3), determine the friction factor<br />

for flow at 2.5 m/s in a 63 mm ID pi<strong>pe</strong>, using the Darby method. Determine also the<br />

pressure drop <strong>pe</strong>r unit length.<br />

Solution SI Units<br />

Reynolds number:<br />

Re = = = 6914.6<br />

Hedstrom number:<br />

(0.063)<br />

He = = = 106,249<br />

2 D × 1440 × 20<br />

���<br />

2 �VDI 1440 × 2.5 × 0.063<br />

� ��<br />

� 0.0328<br />

��0 �2 �<br />

(0.0328) 2<br />

m = 1.7 + 40,000/Re<br />

m = 1.7 + 40,000/6914.6 = 7.49<br />

a = –1.47[1 + 0.146 exp(–2.9 × 10 –5He)] = –1.479<br />

fT = 10aRe –0.193<br />

fT = 10 –1.479 × 6914.6 –0.193 = 0.006<br />

fL = �1 + – � � �1 + Re 16 Re<br />

� �<br />

Re 6He �<br />

4<br />

16 Re<br />

� � �3 7<br />

Re 6He 3f LHe fL = �1 + � = 0.00234<br />

m m 1/m<br />

fn = ( f L + f T )<br />

Therefore,<br />

fn = (0.002347.49 + 0.0067.49 ) 1/7.49 16 6914.6<br />

� ��<br />

6914.6 6 × 106,249<br />

= 0.006<br />

is the fanning friction factor


Darcy factor:<br />

fD = 4fN = 0.006 × 4 = 0.024<br />

Pressure drop <strong>pe</strong>r unit length:<br />

dP/dz = �fDV 2 /(2Di) dP/dz = 1440 × 0.024 × 2.52 /(2 × 0.063) = 1,714 Pa/m (0.816 psi/ft)<br />

5-2 FRICTION LOSSES FOR<br />

PSEUDOPLASTICS<br />

Pseudoplastic rheology was extensively discussed in Chapter 3, Section 3.4-2. Examples<br />

of pseudoplastics are listed in Table 3-10.<br />

5-2-1 Laminar Flow<br />

A number of models have been develo<strong>pe</strong>d for pseudoplastic flows. These treat the fluid as<br />

a continuum.<br />

5-2-1-1 The Rabinowitsch–Mooney Relations<br />

Herzog and Weissenburg (1928) develo<strong>pe</strong>d an equation for laminar time-inde<strong>pe</strong>ndent,<br />

viscous non-Newtonian flows. It was subject to further refinements by Rabinowitsch<br />

(1929) and Mooney (1931).<br />

For a circular pi<strong>pe</strong>, a relationship is established between the shear stress � and the absolute<br />

value of the rate of shear � = –du/dr<br />

f(�) = –du/dr<br />

Rabinowitsch and Mooney derived a general relationship for the shear rate at the wall:<br />

du 8V 1 + 3�<br />

–���w = ���� (5-16)<br />

dr DI 4�<br />

where<br />

HOMOGENEOUS FLOWS OF NONSETTLING SLURRIES<br />

d[ln(Di �P/4L)]<br />

� = ��<br />

(5-17)<br />

d[ln(8V/D)]<br />

5-2-1-2 The Metzner and Reed Approach<br />

Metzner and Reed (1955) develo<strong>pe</strong>d an equation for the Reynolds number in laminar<br />

flow as<br />

D I � V 2–� �<br />

where � is defined by Equation 5-16 and<br />

� = K�8 (�–1) in SI units<br />

ReMR = � (5-18)<br />

�<br />

� = g cK�8 (�–1) in USCS units<br />

5.11


5.12 CHAPTER FIVE<br />

K� = K� � n 1 + 3n<br />

�<br />

4n<br />

In SI units (5-19a)<br />

1 + 3n n<br />

K� = K\gc�� 4n � In USCS units (5-19b)<br />

The fanning friction factor is then expressed in the laminar flow regime in the conventional<br />

manner but using the modified Reynolds number:<br />

16<br />

fNL = �<br />

ReMR<br />

(5-20)<br />

Example 5-5<br />

The pressure drop in a 80 mm ID pi<strong>pe</strong> is to be determined for a <strong>slurry</strong> with S.G. = 1.37.<br />

The power law exponent had been previously determined to be 0.4, and the power law<br />

factor K as 16 dynes-spcn /cm2 . The s<strong>pe</strong>ed of the flow is 1.35 m/s. Use the Metzner and<br />

Reed approach to calculate the friction factor, assuming a shear rate of 600 s –1 .<br />

Solution<br />

From Equation 5-15:<br />

8V 1 + 3�<br />

–600 = ��� 4� �<br />

Di<br />

8 × 1.35 1 + 3�<br />

–600 = ��� 0.08 4� �<br />

–4.45 = (1 + 3�)/4�<br />

–1.11 = 1/� + 3<br />

Re MR =<br />

K� = K� � n<br />

� = 0.529<br />

= 16� � n 1 + 1.2<br />

� = 18.174<br />

1.6<br />

� = K�8 �–1 = 18.174 × 8 –0.47 = 6.84<br />

Re MR =<br />

1 + 3n<br />

�<br />

4n<br />

D � V 2–� �<br />

� �<br />

0.08 0.529 × 1.35 1.471 × 1370<br />

���<br />

6.84<br />

ReMR = 81.87<br />

16<br />

fn = � = 0.195<br />

ReMR<br />

The Metzner and Reed approach has become a classical method of dealing with<br />

time-inde<strong>pe</strong>ndent non-Newtonian fluids. It has been extended to Bingham slurries but<br />

the opinion of the author is that this approach is fairly difficult to use for Bingham slur-


ies, and a more practical method would be the Darby approach, described in Section 5-<br />

1-4.<br />

The Metzner and Reed approach requires the engineer to assume a value of the shear<br />

stress at the wall � w to calculate x. Such assumptions are very difficult to make for engineers<br />

outside a research lab. It may be more practical to send samples of the <strong>slurry</strong> to a<br />

rheology lab and to go from plots of yield stress versus weight concentration, as well as<br />

from plots of viscosity versus weight concentration and shear rates to a more straightforward<br />

computation of friction factors (Figure 5-3).<br />

5.2.1.3 The Tomita Method<br />

Tomita (1959) defined a fanning friction factor for power law fluids as<br />

In the laminar flow regimes:<br />

HOMOGENEOUS FLOWS OF NONSETTLING SLURRIES<br />

5.13<br />

2Di�P 1 + 2n<br />

fPL = ���� (5-21)<br />

3L�mV 1 + 3n<br />

2<br />

n 2–n [1/n + 3] D i V �m<br />

RePL = � ����� (5-22)<br />

K<br />

1–n<br />

6<br />

� ��<br />

n 2 1/n + 2<br />

FIGURE 5-3 Friction factor versus Reynolds number for power law factors. [From Hill R.<br />

A, P. E. Snoek, and R. L. Ghandi, Hydraulic transport of solids, in The Pump Handbook, 2nd<br />

ed., I. J. Karassik et al. (Eds.), New York: McGraw-Hill. Reproduced by <strong>pe</strong>rmission.]


5.14 CHAPTER FIVE<br />

16<br />

fPL = � (5-23)<br />

Remod<br />

5-2-1-3 Heywood Method<br />

Heywood (1991) proposed to define a modified Reynolds number for pseudoplastics as<br />

Remod = � � n<br />

� � n–1<br />

(5-24)<br />

where K and n are the consistency coefficient and flow behavior indexes for pseudoplastic<br />

flows previously defined in Chapter 3.<br />

In the laminar flow regime, Heywood (1991) used the conventional method of defining<br />

the fanning friction factor in terms of the Reynolds number in the laminar flow regime<br />

as previously discussed in Chapter 2, or<br />

fNPL = 16/Remod (5-25)<br />

The effective pi<strong>pe</strong>line viscosity is expressed as<br />

�e = K� � n<br />

� � n–1<br />

�mVDi 4n Di � � �<br />

K 1 + 3n 8V<br />

4n 8V<br />

� � (5-26)<br />

1 + 3n Di<br />

5-2-2 Transition Flow Regime<br />

Ryan and Johnson (1959) defined a critical Reynolds number for purely viscous pseudoplastics<br />

as<br />

Rec = (5-27)<br />

The friction factor at the transition from laminar to turbulent, flow called the critical friction<br />

factor is<br />

(1 + 3n) 1<br />

fNc = ��<br />

(5-28)<br />

(n+2)/(n+1)<br />

(n + 2)<br />

Table 5-4 tabulates the critical Reynolds number and fanning friction factor versus the<br />

power factor “n.” The minimum friction factor is 0.0067 at n = 0.5. However, Heywood<br />

(1991) deducted from various test data that the minimum value for fNc = 0.004, which is<br />

even lower than the values indicated by Equation 5-28 (Figure 5.4).<br />

2<br />

6464n (n + 2)<br />

�<br />

404n<br />

(n+2)/(n+1)<br />

���<br />

(1 + 3n) 2<br />

5-2-3 Turbulent Flow<br />

Various equations have been develo<strong>pe</strong>d over the years for turbulent flow of pseudoplastics<br />

in smooth pi<strong>pe</strong>s. These equations are based on empirical data and semitheoretical<br />

models.<br />

Using the modified Reynolds number as <strong>pe</strong>r Equation 5-17, Dodge and Metzner<br />

(1959) develo<strong>pe</strong>d the following semitheoretical equation for turbulent flow:


HOMOGENEOUS FLOWS OF NONSETTLING SLURRIES<br />

TABLE 5-4 Critical Reynolds Number and Fanning Friction Factor versus the Flow<br />

Behavior Index “n” According to the Ryan and Johnson Method<br />

5.15<br />

Flow Critical Critical Flow Critical Critical<br />

behavior Reynolds fanning friction behavior Reynolds fanning friction<br />

index “n” number factor, f NC index “n” number factor, f NC<br />

0.1 1577 0.01015 0.9 2158 0.00741<br />

0.2 2143 0.00747 1.0 2099 0.00762<br />

0.3 2345 0.00682 1.1 2043 0.0783<br />

0.4 2396 0.00668 1.2 1990 0.00804<br />

0.5 2381 0.00672 1.3 1941 0.00824<br />

0.6 2337 0.00685 1.4 1895 0.00844<br />

0.7 2280 0.00702 1.5 1852 0.00864<br />

0.8 2219 0.0072 1.6 1812 0.00883<br />

1 4<br />

(1–� /2) 0.4<br />

� = � log10[Remod f NT ] – � (5-29)<br />

0.75<br />

1.2<br />

�fN� T� � �<br />

Although Equation 5-29 has been extensively used, it has its own limitations. Measuring<br />

the power exponent “n” in laminar flow tests and then trying to apply it to turbulent<br />

flows is asking for trouble, particularly for cases when n < 0.5. Heywood and Richardson<br />

(1978) showed that pumping flocculated clays yielded higher ex<strong>pe</strong>rimental values of friction<br />

coefficient than those predicted by Dodge and Metzner (1959), particularly when the<br />

value of “n” had been obtained at low shear stress.<br />

Note: Equation 5-20 does not incorporate the effects of roughness. Govier and Aziz<br />

(1972) indicated that Equation 5-20 gives excellent agreement between calculated and ex<strong>pe</strong>rimental<br />

data in the range of modified Reynolds numbers ReMR of 2900–36,000 and<br />

Critical fanning friction factor<br />

( from equation 5-28)<br />

0.010<br />

0.008<br />

0.006<br />

0.004<br />

0.002<br />

f NCR<br />

0.0<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6<br />

Flow index "n"<br />

Re CR<br />

2500<br />

2300<br />

2100<br />

1900<br />

1700<br />

1500<br />

FIGURE 5-4 Values of critical Reynolds number and critical fanning factor versus the flow<br />

index “n” for pseudoplastic slurries based on Equation 5-27.<br />

Modified nritical Reynolds number<br />

(from Equation 5-27)


5.16 CHAPTER FIVE<br />

modified power exponent s of 0.36–1.0. Equation 5-20 requires re<strong>pe</strong>ated iteration. The<br />

use of a <strong>pe</strong>rsonal computer is recommended.<br />

For power law fluids, Tomita (1959) extended his laminar flow model (discussed in<br />

Section 5.2.1.3) to turbulent flows in smooth pi<strong>pe</strong>s by applying Prandtl’s mixing length<br />

concept, and develo<strong>pe</strong>d a different implicit equation:<br />

1<br />

� = 4 log10(RePL �fP� LT �) – 0.40 (5-30)<br />

�fP� �LT<br />

where RePL and fPLT were already expressed for the laminar flow in Equations 5-21 and 5-<br />

22. Tomita’s equation was supported by 40 ex<strong>pe</strong>rimental data points on starch pastes and<br />

lime slurries.<br />

Irvine (1988) published the following equation:<br />

F�(n)<br />

fn = ��<br />

(5-31)<br />

where<br />

F�(n) = � � 1/(3n+1)<br />

2<br />

� ��<br />

n–1 7n n<br />

8 7 (1 + 3n)<br />

(5-32)<br />

Example 5-6<br />

At a volumetric concentration of 30%, a magnetite sus<strong>pe</strong>nsion has a power law coefficient<br />

K of 12 dynes-secn /cm2 and a power law exponent of 0.2. If the <strong>slurry</strong> is homogeneous<br />

and nonsettling at a s<strong>pe</strong>ed of 1.5 m/s, determine the friction factor in a 101 mm ID<br />

pi<strong>pe</strong>, at a <strong>slurry</strong> density of 1600 kg/m3 .<br />

Solution<br />

From Equation 5-24, the modified Reynolds number is calculated as<br />

Remod = � � 0.2<br />

� � –0.8<br />

1600 × 1.5 × 0.101 4 × 0.2 0.101<br />

�� �� � = 202 × 0.5 × 45,697<br />

12 × 0.1 1 + 3 × 0.2 8 × 1.5<br />

Remod = 4615<br />

It is necessary to check if the flow is turbulent.<br />

From Equation 5-27:<br />

Re c� =<br />

[1/(3n+1)]<br />

Remod<br />

6464 n(n + 2) (n+2)/(n+1)<br />

���<br />

(1 + 3n) 2<br />

Rec� = = 2143<br />

Since Remod > Rec�, flow is therefore turbulent. Two different approaches will be used.<br />

Irvine Method<br />

From Equation 5-32:<br />

F�(n) = � � 1/1.6<br />

8 × 0.2<br />

= 0.862<br />

From Equation 5-31, the fanning friction factor is<br />

0.2<br />

6464 × 0.2(2.2)<br />

2<br />

� ��<br />

–0.8 1.4 0.2<br />

8 7 (1 + 0.6) (2.2/1.2)<br />

���<br />

(1 + 0.6) 2<br />

8n n


0.862<br />

fn = � = 0.00442<br />

1/1.6 4615<br />

Tomita Method<br />

From Equation 5-22:<br />

RePL = 4615 × 80.8 × 20.2 × 6(1.6/0.2) 0.8 /2.8 = 316,457<br />

From Equation 5-30:<br />

5.3 FRICTION LOSSES FOR<br />

YIELD PSEUDOPLASTICS<br />

Yield pseudoplastics were described extensively in Chapter 3. Examples were listed in<br />

Table 3.11.<br />

5-3-1 The Hanks and Ricks Method<br />

In the laminar flow regime, Hanks and Ricks (1978), defined the fanning friction factor in<br />

terms of the modified Reynolds number:<br />

fNPL =<br />

16<br />

(5-33)<br />

where<br />

HOMOGENEOUS FLOWS OF NONSETTLING SLURRIES<br />

1<br />

� = 4 log10(316,457 × �fP� LT �) – 0.40<br />

�fP� �LT<br />

Iteration 1 starts by assuming at fPLT � 0.004<br />

correction fPLT = 0.0035<br />

Iteration 2 starts at fPLT � 0.0042<br />

fPLT = 0.00352<br />

Iteration 3 starts at fPLT � 0.0045<br />

fPLT = 0.003498<br />

Iteration 4 starts at fPLT � 0.0035<br />

fPLT = 0.00359<br />

So by theTomita method we obtain a friction factor of 0.0035.<br />

The Irvine method yields a friction factor 23% higher than the Tomita method. Both<br />

methods do not account for roughness of the pi<strong>pe</strong> wall.<br />

� �Remod<br />

� = (1 + 3n) n (1 – x) 1+n� + + � n x2 (1 – x) 2x(1 – x)<br />

� �<br />

1 + 2n 1 + n<br />

2<br />

�<br />

1 + 3n<br />

2 �yp x = = ��<br />

fN · � · V<br />

where �yp is the yield stress for pseudoplastic.<br />

2<br />

�yp �<br />

�w<br />

5.17<br />

(5-34)


5.18 CHAPTER FIVE<br />

For yield-pseudoplastics, Hanks and Ricks (1978), Heywood (1991) proposed to define<br />

a modified Hedstrom Number as<br />

Hemod = � � 2–n/n<br />

2 Di �m �yp � � (5-35)<br />

K K<br />

The critical Reynolds number is established in terms of the modified Reynolds number<br />

and Hedstrom number, as in Figure 5-5.<br />

5-3-2 The Heywood Method<br />

For the laminar flow regime fanning friction factor, Heywood (1991) modified the Buckingham<br />

equation to<br />

fNLY = �1 + 2Hemod ��1 – ���<br />

(2n + 1) fNLY(Remod) 2<br />

2Hemod ��<br />

fNLY(Remod) 2<br />

16<br />

�<br />

Remod<br />

�1 + �1 + 2nHemod �����<br />

(5-36)<br />

fNLY(Remod) 2<br />

4nHemod ���<br />

(n + 1) fNLY(Remod) 2<br />

5-3-3 The Torrance Method<br />

Defining x = � y/� w, Torrance (1963) derived the following equation for turbulent friction<br />

factor of a yield pseudoplastic:<br />

Critical (Hanks & Ricks) Reynolds Number<br />

3200<br />

2800<br />

2400<br />

2000<br />

1600<br />

1200<br />

800<br />

400<br />

0<br />

0<br />

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0<br />

Flow Behavior Index "n"<br />

4<br />

He = 10<br />

He = 0<br />

He = 10 6<br />

FIGURE 5-5 Values of the critical value of the “Hanks and Ricks Reynolds number” versus<br />

the flow index “n” for yield pseudoplastic slurries.


� log10RePLC( f [1–n/2] 4.53<br />

)�<br />

+ (5n – 8) (5-37)<br />

where RePLC = DnV 2–n�/K8n–1 0.68<br />

� �<br />

n<br />

n<br />

.<br />

The work of Torrance was essentially an exercise in algebra, and has not been substantiated<br />

by ex<strong>pe</strong>rimental data. Nevertheless it has been substantially quoted in the absence<br />

of suitable confirming data.<br />

5-4 GENERALIZED METHODS<br />

Various models have been develo<strong>pe</strong>d for complex non-Newtonian flows. The most important<br />

ones are listed, but most were derived from empirical data.<br />

5-4-1 The Herschel–Bulkley Model<br />

Herschel and Bulkley (1928) develo<strong>pe</strong>d a model that has been extensively applied to<br />

sewage sludge, kaolin slurries, and mine tailings:<br />

� = �y + j(�) n (5-38)<br />

The apparent viscosity is<br />

�a = = + j(�) n–1 � �y � � (5-39)<br />

� �<br />

where j is called the Herschel–Bulkley parameter.<br />

5-4-2 The Chilton and Stainsby Method<br />

Chilton and Stainsby (1998) indicated that the accuracy of the Herschel–Bulkley model<br />

deteriorated at high shear rates. However, this may or may not be significant, de<strong>pe</strong>nding<br />

on the application. At high strain rates the model predicts that the viscosity tends to zero,<br />

which is obviously incorrect.<br />

In Chapter 3, Section 3-4-2-2 , the Sisko, Cross, Meter, and Bird rheological models<br />

were presented. Chilton and Stainsby (1998) stressed the limitations of these models and<br />

the shear rates at which they are valid.<br />

In an effort to solve the Rabinowitsch and Mooney equations presented in Equations<br />

5-15 and 5-16, Chilton and Stainsby (1998) proposed to express the pressure drop for a<br />

Herschel–Bulkley fluid as:<br />

where<br />

�P<br />

� L<br />

HOMOGENEOUS FLOWS OF NONSETTLING SLURRIES<br />

= �� � – 2.95� + log 1 2.69<br />

4.53<br />

�� � � 10(1 – x)<br />

�f� n<br />

n<br />

4j<br />

�<br />

D<br />

8V<br />

�<br />

D<br />

= � � n<br />

� � n<br />

� �� � n<br />

3n + 1 1<br />

1<br />

��<br />

� 4n<br />

� 1 – x<br />

x = = 4L� �0 0<br />

� �Di�P �w<br />

1 – ax – bx 2 – cx 3<br />

5.19


5.20 CHAPTER FIVE<br />

b =<br />

1<br />

a = �<br />

(2n + 1)<br />

2n<br />

��<br />

(n + 1)(2n + 1)<br />

c =<br />

For a Bingham <strong>slurry</strong>, j = � or the coefficient of rigidity. For a power law pseudoplastic<br />

fluid, �0 = 0, j = �p or the power law viscosity. For a Newtonian, Hagen–Poiseuille<br />

fluid, �0 = 0, � = 1, j = � or viscosity.<br />

= � � =<br />

Defining an effective viscosity as<br />

� * = j� � n–1<br />

� � n<br />

� �� � n<br />

2n<br />

�P 4� 8V 32�V<br />

� � � �2 L D D D<br />

8V 3n + 1 1<br />

1<br />

� � � ��<br />

(5-40)<br />

2 3<br />

Di 4n 1 – x 1 – ax – bx – cx<br />

Chilton and Stainsby proposed their equation for a generalized Reynolds number as<br />

�VDi ReMR = (5-41)<br />

2<br />

��<br />

(n + 1)(2n + 1)<br />

� �*<br />

Using the value � (defined by Equation 5-16), the authors indicated that<br />

3� + 1<br />

� * = �L�� 4� �<br />

and a modified Reynolds number is defined as<br />

4��VD i<br />

ReMR = ��<br />

(5-42)<br />

�L(3� + 1)<br />

if the wall viscosity could be measured.<br />

The friction factor in the laminar regime is then expressed as<br />

16<br />

fn = � (5-43)<br />

ReMR<br />

In the turbulent regime, for Herschel–Bulkley fluids Chilton and Stainsby derived<br />

Re MR<br />

fn = 0.079� � –0.25<br />

��<br />

2 4 n (1 – x)<br />

Chilton and Stainsby then proposed another modified Reynolds number:<br />

Re MR<br />

(5-44)<br />

R� =<br />

This indicates that Equation 5-44 is a modified Blasius equation (see Chapter 2):<br />

(5-45)<br />

fn = 0.079(R�) –0.25 �� 2 4 n (1 – x)<br />

(5-46)


HOMOGENEOUS FLOWS OF NONSETTLING SLURRIES<br />

Without further proof, they proposed to follow the Prandtl equation of Newtonian fluids<br />

as if it could be applied to non-Newtonian fluids:<br />

f n –0.5 = 4.0 log10(R�f n 0.5 ) – 0.4 (5-47)<br />

Deviations from ex<strong>pe</strong>rimental data were noticed at high Reynolds numbers due to the<br />

limitations of the Herschel–Bulkley model on which basis this model was develo<strong>pe</strong>d.<br />

(See Figures 5-6 and 5-7.)<br />

Example 5-7<br />

Lab tests are conducted on sewage sludge in a 6� pi<strong>pe</strong>. The density is 1018 kg/m 3 (S.G. =<br />

1.018). The yield stress is measured as 1.28 Pa. The <strong>slurry</strong> is a Herschel–Bulkley fluid<br />

mixture with a parameter j = 0.2. The power law coefficient is determined to be 0.74. Calculate<br />

the friction factor for a flow of 350 L/s in a 18� pi<strong>pe</strong> of 0.375� thickness, assuming<br />

a wall shear stress of 1.6 Pa.<br />

Pi<strong>pe</strong> inner diameter = (18 – 2 × 0.375) × 0.0254 = 0.438 m<br />

Pi<strong>pe</strong> inner area = 0.25 × � × 0.438 2 = 0.151m 2<br />

Pi<strong>pe</strong> inner s<strong>pe</strong>ed = 0.35/0.151 = 2.31 m/s<br />

�0 1.28<br />

x = � = � = 0.8<br />

�w 1.6<br />

5.21<br />

FIGURE 5-6 The friction factor versus the Chilton–Stainsby Reynolds number for Bingham<br />

mixtures. (From A. Chilton and R. Stainsby, 1998. Reproduced with <strong>pe</strong>rmission of Journal of<br />

Hydraulic Engineering, ASCE.)


5.22 CHAPTER FIVE<br />

FIGURE 5-7 The friction factor versus the Chilton–Stainsby Reynolds number for pseudoplastic<br />

mixtures. (From A. Chilton and R. Stainsby, 1998. Reproduced with <strong>pe</strong>rmission of<br />

Journal of Hydraulic Engineering, ASCE.)<br />

� * = 0.2� � (0.74–1)<br />

8 × 2.31 3 × 0.74 + 1<br />

� ��<br />

0.438<br />

4 × 0.74<br />

1<br />

= 0.366���2 3<br />

1 – ax – bx – cx �<br />

� * = 0.2� � –0.26<br />

8 × 2.31<br />

�<br />

0.438<br />

� � 0.7 1<br />

1<br />

� ����� � � 0.75<br />

(2.12 + 1)<br />

��<br />

2.96<br />

1<br />

5 ������<br />

� 1 – 0.403 × 0.064 – 0.343 × 0.064 �<br />

2 – 0.254 × 0.0643 � * = 0.3725<br />

Re MR = 1018 × 2.31 × 0.438/0.3725 = 2765<br />

This is transition Reynolds Number between laminar and turbulent flow. In laminar<br />

regime fn = 16/ReMR = 0.0057.<br />

�fnV Pressure drop = = 72 Pa/m<br />

2 �P 2<br />

� �<br />

L Di<br />

5-4-3 The Wilson–Thomas Method<br />

� 1 – 0.8<br />

1 – ax – bx 2 – cx 3<br />

The Wilson–Thomas Method was develo<strong>pe</strong>d in the 1980s for yield pseudoplastic and<br />

power law slurries. Wilson (1985) and Thomas and Wilson (1987) assumed that the fluid


is a continuum, but that in non-Newtonian cases the viscous sublayer was thicker than<br />

with the water. (See Chapter 2 for the definition of viscous sublayer.)<br />

Defining a velocity VN for a Newtonian fluid at the same wall shear stress �w as for the<br />

flow of a non-Newtonian fluid, a bulk velocity V is defined as<br />

V = VN + Uf�2.5 ln� � + 11.6� �� (5-48)<br />

with<br />

= 2.5 loge� � (5-49)<br />

Since Uf = (�w/�m) 1/2 , the Wilson–Thomas model requires the designer to assume a value<br />

of wall shear stress.<br />

The effective viscosity is defined as<br />

�eff = � � n–1<br />

� � n–1<br />

(5-50)<br />

which is slightly different than expressed by Equation 5-26.<br />

In the laminar flow regime, the shear rate is expressed as<br />

= � � 1/n<br />

(5-51)<br />

For Bingham fluids, the Wilson–Thomas equation is written as<br />

V = VN + Uf�2.5 ln� � + x (14.1 + 1.25x)�<br />

n + 1 1 – n<br />

� �<br />

2<br />

n + 1<br />

VN �Di Uf � �<br />

Uf<br />

�<br />

4n 8V<br />

� �<br />

3n + 1 Di<br />

8V 4n �w � � �<br />

Di 3n + 1 K<br />

1 – x<br />

� (5-52)<br />

1 + x<br />

where<br />

HOMOGENEOUS FLOWS OF NONSETTLING SLURRIES<br />

x = � 0/� w<br />

ReMR = �mVDi(1 – x)/�� In the laminar flow regime<br />

= �1 – x + x4 8V �w 4 1<br />

� � � � � (5-53)<br />

Di �� 3 3<br />

(where �p = plastic viscosity), which is essentially the Buckingham equation. For Bingham<br />

fluids in the laminar regime, if x � 0.5 then<br />

8V�� 4<br />

�w = � + � �0 Di 3<br />

To obtain the transition velocity from laminar to turbulent flow, the following approach<br />

based on the Wilson–Thomas method is recommended. In the turbulent regime, a<br />

Reynolds number based on the friction velocity U f is defined as<br />

Ref = �DIUf ��p<br />

5.23


5.24 CHAPTER FIVE<br />

The Wilson–Thomas velocity is defined by modifying Equation 5.52 to<br />

1 – x<br />

V = Uf�2.5 ln Re + 2.5 ln� � + x (14.1 + 1.25x)�� (5-54)<br />

1 + x<br />

The transition occurs at the intersection of Equations 5-53 and 5-54.<br />

The Wilson–Thomas method was derived from first principles and does not use empirical<br />

correction coefficients. It has proved to be correct for many slurries, but it sometimes<br />

overpredicts losses for the Carson slurries (described in Equation 3-52).<br />

Example 5-10 shows the method of calculation using the Wilson–Thomas equation.<br />

Figure 5-8 compares the Wilson–Thomas method with the Hanks–Dadia friction factor.<br />

Figure 5-9 compares it with ex<strong>pe</strong>rimental data.<br />

5-4-4 The Darby Method: Taking into Account<br />

Particle Distribution<br />

Professor R. Darby (2000) from Texas A&M University recently published a new method<br />

to predict the friction factor of power law non-Newtonian slurries. It is, however, much<br />

closer to the domain of slurries and takes into account such concepts as the drag coefficient,<br />

to which the reader was exposed in Chapter 3.<br />

In a method reminiscent of the work of Wasp for compound heterogeneous flows<br />

(see Chapter 3), Darby stated that the overall pressure drop for a non-Newtonian fluid<br />

is essentially the pressure drop of the liquid phase plus the pressure drop due to the<br />

solids:<br />

�Pm = �Pf + �Ps (5-55)<br />

for each fraction i of solids with an average diameter dpi, and with a volumetric concentration<br />

Cv, the individual pressure drop must be computed. In a first iteration, the pressure<br />

drop for the liquid phase is computed by treating it as homogeneous non-Newtonian liquid<br />

by the various methods described in this chapter. A Froude number for the solid particles<br />

is defined as<br />

FIGURE 5-8 Comparison between the Wilson–Thomas and the Hanks–Dadia models for<br />

friction factor of Bingham slurries. (From A. D. Thomas and K. C. Wilson, 1985. Reproduced<br />

by <strong>pe</strong>rmission of Canadian Journal of Engineering.)


Fr2 V<br />

= = (5-56)<br />

2<br />

�� ��<br />

(�s/�L – 1)d�g (�s/�L – 1)gDi where V t is the terminal velocity of falling particles.<br />

At slip velocity V r between the solids and the liquid phases, with V as the carrier s<strong>pe</strong>ed<br />

of the sus<strong>pe</strong>nsion, a nondimensional pressure drop is defined as<br />

If V r = V f – V s,<br />

HOMOGENEOUS FLOWS OF NONSETTLING SLURRIES<br />

X = � � 2<br />

��s �� �<br />

Cv�L(�s/�L – 1)gL V<br />

W r 2<br />

(5-57)<br />

X = � (5-58)<br />

1 – Wr<br />

where W r = V r/V. For small volume fraction 0 < C v < 0.25, X = X 0.<br />

For other concentrations,<br />

V t 2<br />

V t<br />

5.25<br />

FIGURE 5-9 Comparison between the Wilson–Thomas and ex<strong>pe</strong>rimental data on pressure<br />

losses for a limestone <strong>slurry</strong>. (From A. D. Thomas and K. C. Wilson, 1985. Reproduced by<br />

<strong>pe</strong>rmission of Canadian Journal of Engineering.)<br />

X = X 0 + 0.1F R 2 (Cv – 0.25) (5.59)


5.26 CHAPTER FIVE<br />

Procedures for Newtonian Mixtures:<br />

1. Calculate the Froude number.<br />

2. Determine Vt from the Archimedean number:<br />

Ar =<br />

d p 3 �f g(� s/� L – 1)<br />

��<br />

� L 2<br />

with � f being the liquid viscosity.<br />

If Ar < 15, the terminal velocity can be obtained from the Stokes equation<br />

If Ar > 15, the particle Reynolds number is obtained after rearranging the<br />

Dallavalle equation:<br />

Re p = [(14.42 + 1.827�F�r�) 0.5 – 3.798] 2 = d pV t � L/� L<br />

(5-60)<br />

3. Calculate the Froude number from Equation 5-56.<br />

4. The value of ratio Wr is then calculated from the Molerus diagram (Figure 5-10). The<br />

values of X and �Ps/L are then calculated from Equations 5-59 and 5-60. This step<br />

may be re<strong>pe</strong>ated for each range of particle size and summed up with other particle<br />

sizes to get an overall pressure drop for solids. However for Non-Newtonian mixtures<br />

additional procedures are needed.<br />

FIGURE 5-10 Molerus diagram. (From R. Darby, 2000. Reproduced by <strong>pe</strong>rmission of<br />

Chemical Engineering.)


Terminal Velocity for Power Law Fluids<br />

For a solid particle settling in a non-Newtonian fluid, defining<br />

Z = (5.61)<br />

as a nondimensional power law parameter and<br />

C1 = [(1.88/n) 8 + 34] –1/8<br />

1.33 + 0.37n<br />

��3.7<br />

1 + 0.7n<br />

as another nondimensional power law parameter, Darby expressed the modified Reynolds<br />

number as<br />

with C d the drag coefficient:<br />

HOMOGENEOUS FLOWS OF NONSETTLING SLURRIES<br />

4.8<br />

�1/2 C d – C1<br />

Re PL = Z� � 2<br />

= � (5.62)<br />

K<br />

4g(�m/�L – 1)d� Cd = ��<br />

(5.63)<br />

Friction Factors<br />

The transition from laminar to turbulent flow occurs at a critical Reynolds number:<br />

RePLC = 2100 + 875(1 – n) (5.64)<br />

Defining fL as the fanning factor in the laminar regime, and fT as the fanning factor in the<br />

turbulent regime, Darby derived the following equations:<br />

�<br />

fn = (1 – �) fL + ��<br />

(5.65)<br />

–8 –8 1/8<br />

( f T + f L )<br />

where the laminar flow friction factor is<br />

16<br />

fL = � (5.66)<br />

RePL<br />

0.0682n –1/2<br />

fT = ��<br />

(5.67)<br />

1/(1.87+2.39n)<br />

RePL<br />

At the transition from laminar to turbulent flow, the friction factor is expressed as:<br />

fTR = 1.79 × 10 –4 (0.414+0.757n) exp(–5.24n)RePL (5.68)<br />

and<br />

1<br />

� = � (5.69)<br />

� 1 + 4<br />

� = Re PL – Re PLC<br />

Example 5-8<br />

Slurry is required to flow in a 250 mm ID pi<strong>pe</strong> and has the following characteristics:<br />

�0 = 15 Pa<br />

� = 43 mPa · s<br />

� = 1450 kg/m3 3V t 2<br />

d � n V t 2–n �f<br />

5.27<br />

(5.70)


5.28 CHAPTER FIVE<br />

K = 0.4 Pa · sn n = 0.8<br />

d85 = 65 �m<br />

V = 1.76 m/s<br />

Determine the friction factor. Check on critical Reynolds number:<br />

RePLC = 2100 + 875(1 – 0.8) = 2275<br />

Z = 1.13<br />

C1 = 0.4236<br />

Calculations will be done by a series of iterations.<br />

Iteration 1. Assume the drag coefficient is Cd = 0.4, then<br />

RePL = 1.13� � 2 4.8<br />

�� = 597<br />

1/2 0.4 – 0.4236<br />

16 16<br />

fL = � = � = 0.026<br />

RePL 597<br />

0.0682 × 0.8<br />

fT = = 0.0141<br />

–1/2<br />

��<br />

597 0.264<br />

� = 597 – 2275 = –1678<br />

� =<br />

� � 0<br />

�<br />

fn = (1 – �)fL + ��<br />

–8 –8 1/8<br />

( f T + f L )<br />

fn = fL = 0.026<br />

By further iteration, the magnitude of the friction factor is refined.<br />

5-5 TIME-DEPENDENT<br />

NON-NEWTONIAN SLURRIES<br />

1<br />

� 1 + 4 –�<br />

Time-de<strong>pe</strong>ndent non-Newtonian flows are difficult to model. There is a certain lapse of<br />

time to overcome before a stable friction factor can be obtained. Govier and Aziz (1972)<br />

have published cases in which the lapsing time was as high as 1000 minutes with Prembina<br />

crude oil. During this initial lapse of time, the pressure gradient for friction losses<br />

drop<strong>pe</strong>d from an initial value of 72.5 Pa/m (0.0032 psi/ft) down to 18 Pa/m (0.0008 psi/ft)<br />

by the time the flow had stabilized. Such a ratio of 4:1 is certainly not negligible.<br />

Govier and Aziz (1972) indicated that once the initial <strong>pe</strong>riod of stabilization is<br />

reached, the general form of pressure loss equations are the same as for time-inde<strong>pe</strong>ndent<br />

non-Newtonian fluids. At the entry to a pi<strong>pe</strong>, the flow may be laminar, but at a certain distance,<br />

once the entrance effects are overcome, the flow can transit to turbulence.<br />

The start-up pressures for thixotropic slurries may be quite high, particularly when<br />

these slurries coagulate into gels inside the pi<strong>pe</strong>line. During an initial of <strong>pe</strong>riod of time,<br />

gels must be ex<strong>pe</strong>lled from the pi<strong>pe</strong>line. Crude oils, fuel oils, bentonite clay, and drilling


muds are a concern to engineers. Water may be used as an ex<strong>pe</strong>lling medium to clear up<br />

the pi<strong>pe</strong>line. Positive displacement pumps are preferred for thixotropic slurries to overcome<br />

the high starting pressures.<br />

Various models have been develo<strong>pe</strong>d for thixotropic fluids. These slurries are sometimes<br />

treated as Bingham plastics and sometimes as pseudoplastics, based on the ex<strong>pe</strong>rimental<br />

data from test work.<br />

The problem of pumping thixotropic froth in oil sand plants was a challenge to manufacturers<br />

of pumps, and for a while the belief was that only positive displacement pumps<br />

could be used. In 2000, the o<strong>pe</strong>rators of oil-sand processing plants invited the manufacturers<br />

of <strong>slurry</strong> pumps to develop new appropriate pumps. Research is being conducted at<br />

the Saskatchewan Research Institute in Canada and is starting to yield new concepts of<br />

pump design.<br />

5-6 EMULSIONS<br />

HOMOGENEOUS FLOWS OF NONSETTLING SLURRIES<br />

The concept of emulsions was introduced in Chapter 1. In an emulsion, one phase is sus<strong>pe</strong>nded<br />

as droplets rather than particles. Tar or bitumen, for example, can be sus<strong>pe</strong>nded in<br />

water up to a volumetric concentration of 70%. In some of the oil sand extraction processes<br />

in Canada, the situation can be further complicated by the addition of flocculants. The<br />

Venezuelan corporation, PDVSA, and its research branch, INTEVEP, conducted considerable<br />

research on Orimulsion, a proprietary synthetic fluid composed of surfactants,<br />

bitumen droplets, and water. The surfactants keep the bitumen droplets in sus<strong>pe</strong>nsion.<br />

Nunez et al. (1996) published a comprehensive pa<strong>pe</strong>r on the flow characteristics of<br />

concentrated emulsions in water with a volumetric concentration in the range of 70–85%.<br />

5-7 ROUGHNESS EFFECTS ON FRICTION<br />

COEFFICIENTS<br />

5.29<br />

Szilas et al. (1981) published the following equation for pseudoplastic oil flow in rough<br />

pi<strong>pe</strong>s at high Reynolds numbers:<br />

= log10 (Remod(4fn) 1–n/2 ) + 1.511/n�4.24 + � – – 2.114 (5-71)<br />

However, this equation does not include any terms for wall roughness.<br />

Torrance (1963) develo<strong>pe</strong>d a theoretical equation for yield pseudoplastics in fully turbulent<br />

flow (at high Reynolds numbers) in rough pi<strong>pe</strong>s as<br />

= 4.07 log10� � + 6.0 – (5-72)<br />

where RePLC = �DnV 2–n /K8n–1 1 4<br />

1.414 8.03<br />

� � � �<br />

�fn� n<br />

n n<br />

1<br />

RePLC 2.65<br />

� � �<br />

�fn�<br />

�<br />

n<br />

. This applies to the region where the friction factor is inde<strong>pe</strong>ndent<br />

of the Reynolds number, or at high Reynolds numbers.<br />

Govier and Aziz (1972) suggest the following procedure:<br />

� Compute the friction factor for the <strong>slurry</strong> in a smooth tube using one of the methods<br />

described in Sections 5-2, 5-3, and 5-4.<br />

� Using the Moody diagram determine the ratio of the friction factor for rough to smooth<br />

pi<strong>pe</strong> at the value of the Reynolds number (using Re B, Re PLC, or Re mod).


5.30 CHAPTER FIVE<br />

Aral and Kaylon (1994) focused on highly concentrated sus<strong>pe</strong>nsions and investigated the<br />

effects of tem<strong>pe</strong>rature as well as surface roughness.<br />

To take into account the pi<strong>pe</strong> roughness in laminar and turbulent flows, SRC (2000)<br />

recommended the use of the equation of Churchill (1977):<br />

where<br />

8<br />

�<br />

Re<br />

f n = 2�� � 12<br />

+ (A + B) –1.5 � 1/12<br />

A = � –2.457 ln�� � 0.9 7<br />

16<br />

� + 0.27 �/D<br />

Re<br />

i��<br />

B = � � 16 37,530<br />

�<br />

Re<br />

(5-73)<br />

(5.74)<br />

(5.75)<br />

Example 5-9<br />

Assuming turbulent flow and using the Wilson–Thomas model, calculate the bulk velocity<br />

for a flow with the following characteristics:<br />

�0 = 20 Pa<br />

�w = 24 Pa<br />

�� = 0.016 Pa · s<br />

� = 3 × 10 –6 m<br />

Di = 141 mm<br />

� = 1350 kg/m3 x = �/�w = 20/24 = 0.833<br />

Iteration 1. Assume VN = 2.5 m/s, then<br />

Re = �VDi/� � = ��/(1 – x) = 0.016/(1 – 0.833) = 0.096<br />

Re = �VDi/� Re = 1350 × 2.5 × 0.141/0.096 = 4967<br />

Relative roughness �/Di = 3 × 10 –6 /0.141 = 0.000021<br />

From Equations 5-74 and 5-75:<br />

A = {–2.457 ln[(7/4967) ) 0.9 + 0.27 × 0.000021)] 16 = 3.86 × 1018 B = 1.1286 × 10 14<br />

f n = 0.00949<br />

Iteration 2.<br />

Checking the value of V N. From Chapter 2:<br />

�w = fn�V 2 /2<br />

(2 × 24)/(0.00949 × 1350) = 3.75<br />

V = 1.935 m/s<br />

The friction velocity from Equation 2-5:<br />

Uf = ��w�/�� = 0.133 m/s


HOMOGENEOUS FLOWS OF NONSETTLING SLURRIES<br />

Using Equation 5-52:<br />

1 – 0.833<br />

V = 1.935 + 0.133�2.5 ln � + 0.833(14.1 + 1.25 × 0.833)�<br />

1.833<br />

2.815 m/s = 1.935 + 0.133[–5.989 + 12.61]<br />

Therefore, the equivalent Newtonian fluid velocity is 1.935 m/s, and the mean velocity of<br />

the Bingham <strong>slurry</strong> will effectively be 2.815 m/s.<br />

Up until 1990, one author after another tried to treat non-Newtonian slurries as homogeneous<br />

mixtures. The arbitrary assumption that the particle diameter was absent in many<br />

equations when the diameter was smaller than 44 �m or 74 �m (de<strong>pe</strong>nding on the author)<br />

is now being challenged. The advent of new ex<strong>pe</strong>rimental methods, such as laser velocimeters,<br />

is helping engineers understand the complexity of turbulent non-Newtonian<br />

fluids. The work of Park et al. (1989) using laser velocimeters, and Pokryvalio and<br />

Grozberg (1995) using electro-diffusion techniques on non-Newtonian slurries was analyzed<br />

by Slatter et al. (1996), who confirmed the significance of high-intensity turbulence<br />

at the wall. This encouraged them to postulate a theory of particle roughness. Slatter et al.<br />

proposed that the d 85 particle size be used for particle roughness. (This may be an attempt<br />

to correlate with the Nikrudase particle roughness of Newtonian slurries.) For large pi<strong>pe</strong>,<br />

when the actual roughness � exceeds the value d 85, � is used. For<br />

� < d 85<br />

� > d 85<br />

d x = d 85<br />

d x = �<br />

For a yield pseudoplastic, Slatter et al. (1996) defined their Reynolds number, Re r, in<br />

terms of the friction velocity, consistency factor K, and power coefficient n, as well as<br />

roughness d x:<br />

8�U f 2<br />

5.31<br />

Rer = (5.76)<br />

If Rer > 3.32, then smooth wall turbulence occurs, and the mean bulk velocity V is expressed<br />

as<br />

= 2.5 ln� � – 2.5 ln Re ��<br />

�0 + K(8Uf /dx) V Di � � r + 1.75 (5.77)<br />

Uf 2dx<br />

If Rer ( 3.32, then fully develo<strong>pe</strong>d rough turbulent flow occurs, and the mean bulk velocity<br />

V is expressed as<br />

V Di � = 2.5 ln��� + 4.75 (5.78)<br />

Uf 2dx<br />

This correlation produces an abrupt transition from smooth turbulent to fully turbulent<br />

flow at the wall of the pi<strong>pe</strong>.<br />

We have already explained that the Wilson–Thomas model was based on the assumption<br />

that the viscous sublayer in non-Newtonian flows was thicker than with water, thus<br />

suppressing the effect of roughness. The work of Slatter, Thorscalden, and Petersen<br />

(1996) on mixtures of kaolin clay and sand indicates that this is not very achievable (Figure<br />

5-11). The resultant pressure losses (Figure 5-12) are therefore higher. Figure 5-12<br />

does indicate that the Torrance and the Wilson–Thomas models correlate well. Both models<br />

were based on mathematical assumptions at 20 year intervals.<br />

n


5.32 CHAPTER FIVE<br />

FIGURE 5-11 A comparison between the Wilson–Thomas and Slatter, Thorscalden, and<br />

Petersen models for the viscous sublayer. (From P. T. Slatter et al., 1996. Reproduced by <strong>pe</strong>rmission<br />

of BHR Group.)<br />

FIGURE 5-12 A comparison of pressure drop <strong>pe</strong>r unit length between the Slatter,<br />

Thorscalden, and Petersen and Wilson–Thomas models. (From P. T. Slatter et al., 1996. Reproduced<br />

by <strong>pe</strong>rmission of BHR Group.)


Example 5-10<br />

A sand–kaolin mixture with a volumetric concentration of 19.5% is flowing in a 431 mm<br />

ID pi<strong>pe</strong> at an average s<strong>pe</strong>ed of 1.8 m/s. The yield stress is 5.5 Pa, the coefficient of consistency<br />

K = 0.124 Pa · sn , and the power coefficient n = 0.64. The s<strong>pe</strong>cific gravity of the solids<br />

are 2.65 and the d85 = 131 �m. Using the Slatter method, determine the friction factor.<br />

Solution<br />

Density of mixture:<br />

�m = Cv(�s – �L) + �L �m = 0.19(1265) + 1000 = 1240 kg/m3 Iteration 1. Assume fully turbulent flow:<br />

= 2.5 ln� � = 4.75 = 23.26<br />

If V = 1.8 m/s, then Uf = 0.0774 m/s, since Uf = V �fD�/8�. fD = 0.0147<br />

8 × 1240 × 0.0774<br />

Rer = = 1.78<br />

Iteration 2. Since Rer < 3.32, the equation to use is<br />

2<br />

�����<br />

5.5 + 0.124(8 × 0.0774/0.131 × 10 –3 ) 0.64<br />

V<br />

0.5Di � ��–6<br />

Uf 131 × 10<br />

V<br />

� Uf<br />

5-8 WALL SLIPPAGE<br />

HOMOGENEOUS FLOWS OF NONSETTLING SLURRIES<br />

0.5D i<br />

= 2.5 ln� � + 2.5 ln Re � r + 1.75<br />

d85<br />

V<br />

� = 18.51 + 1.44 + 1.75<br />

Uf<br />

V<br />

� = 21.70<br />

Uf<br />

0.046 = (fD/8) fD = 0.01698<br />

Uf = 0.0829 m/s<br />

5.33<br />

A phenomenon encountered with non-Newtonian mixtures is a tendency for the low-viscosity<br />

constituent to migrate to regions of high shear and to lubricate the flow. One example<br />

is the core annular flow of crude oil in water, where the more viscous material is lubricated<br />

by the less viscous material. In the case of emulsions and certain non-Newtonian<br />

slurries, lubrication occurs by a slip layer of water on the wall.<br />

Mathematically, the concept of slip can be treated as a discontinuity. Heywood (1991)<br />

proposed to represent slip by the slip velocity Vs. In the laminar flow regime, the total<br />

flow in a pi<strong>pe</strong> would be<br />

Q = Vs + � � 3<br />

� �w �<br />

0<br />

2 2 �D i � Di d�<br />

� � � � d� (5-79)<br />

4 8 �w dt


5.34 CHAPTER FIVE<br />

Heywood (1991) noticed that there are no methods to evaluate slip in turbulent flows.<br />

Evaluation of slip in laminar flow is conducted using the coaxial cylinder described in<br />

Chapter 3.<br />

Nunez et al. (1996) indicate that the migration of viscous droplets from the wall in an<br />

emulsion exhibit the Segré–Silberberg effect. However, they pointed out that not all<br />

emulsions ex<strong>pe</strong>rience slip and that the phenomena of slip ap<strong>pe</strong>ared to be characteristic of<br />

the crude oil or viscous component. In some res<strong>pe</strong>cts, slip does reduce the friction factor<br />

by lubrication from the least viscous phase. However, Aral and Kaylon (1994) found that<br />

increasing the surface roughness tended to reduce or eliminate slip.<br />

One particular problem with emulsions is the fracture of the droplets under high<br />

shear rates. A form of comminution occurs as large droplets come in contact with other<br />

ones. Degradation of non-Newtonian slurries under high shear rates is not well documented.<br />

There is evidence that high shear rates occur in centrifugal pumps. Adequate clearance<br />

may reduce the degradation of the emulsion or <strong>slurry</strong> but tends to reduce the efficiency of<br />

the pump. Degradation can include a form of coalescence or formation of colloids and<br />

larger droplets or particles. Clay ball formation is encountered in dredging o<strong>pe</strong>rations after<br />

passage through the pump.<br />

5-9 PRESSURE LOSS THROUGH<br />

PIPE FITTINGS<br />

The method of the two K-factor was presented in Chapter 2 in Section 2.8. It was explained<br />

that Hoo<strong>pe</strong>r had established a general relationship<br />

K1 1<br />

K = � + K��1 + �<br />

Re � DI-in<br />

where<br />

K1 is the value of K at a Reynolds number of 1<br />

K� is the value of K at high Reynolds number<br />

Di-in is the internal pi<strong>pe</strong> diameter in inches<br />

Johnson (1982) reviewed some of the problems of pumping non-Newtonian mixtures.<br />

In his assessment of fittings for sewage, he indicated important discrepancies in the laminar<br />

regime with losses 2–4 times as much as those for water flows. In the turbulent<br />

regime, the losses were either of the order of those for water or higher. He recommended<br />

that further studies be conducted for laminar flow, but for turbulent flows, the concept of<br />

equivalent length be used.<br />

In Chapter 2, concepts of pressure losses for Newtonian liquids were examined. Edwards<br />

et al. (1985) reviewed the flow of non-Newtonian slurries in laminar regimes. They<br />

recommended that the modified Reynolds number (using ReB or Remod) be used to correlate<br />

with the loss factor Kf of Newtonian flows in laminar regimes.<br />

In certain fittings such as globe valves, turbulence is enhanced by the geometry of the<br />

fitting. Although the flow may be laminar in a straight pi<strong>pe</strong> up to a transition to a<br />

Reynolds number of 2000, turbulence in the globe valve may actually start at much lower<br />

Reynolds number such as 900.<br />

Much more work is needed on loss factors for non-Newtonian flows in transition and<br />

turbulent regimes. Govier and Aziz (1972) had noticed the lack of any methodology to<br />

compute loss from pi<strong>pe</strong> fittings with non-Newtonian flows. They proposed that the


method of equivalent length be used, i.e., that the equivalent length of pi<strong>pe</strong> fittings for<br />

Newtonian liquids be added to computations of total length for pressure loss of the non-<br />

Newtonian <strong>slurry</strong>.<br />

5-10 SCALING UP FROM SMALL<br />

TO LARGE PIPES<br />

Non-Newtonian flows are complex. A lab test in a pumping loop can yield very useful<br />

rheology data about the pressure drop. Scaling up to larger pi<strong>pe</strong>s is one method of predicting<br />

pi<strong>pe</strong>line flows. Heywood et al. (1992) proposed the following methods:<br />

� For laminar flow, the plot of the shear rate (8V/Di) against the shear stress (Di�P/4L) is<br />

inde<strong>pe</strong>ndent of the pi<strong>pe</strong> diameter and the pi<strong>pe</strong> roughness, so that ex<strong>pe</strong>rimental data<br />

could be converted directly into practical data for pi<strong>pe</strong>line design.<br />

� For turbulent flows, the Bowen method, which is essentially a modification of the Blasius<br />

method, should be used. Bowen (1961) suggested the following modification to<br />

the Blasius equation:<br />

x w Di � = kV (5.80)<br />

The shear stress is plotted against the flow rate Q to obtain the magnitude of the exponent<br />

w. The result is then used to plot �/V w against the diameter Di to obtain the values of k and<br />

x. The intersection of the turbulent and laminar flow curves gives the transition point.<br />

Kenchington (1972) showed that this method showed great discrepancy when the ratio of<br />

diameter between the large field pi<strong>pe</strong> and the lab pi<strong>pe</strong> exceeded 6 to 12 folds, so that large<br />

pi<strong>pe</strong> tests may be needed. The Bowen method assumed the same range of roughness between<br />

a lab and a field pi<strong>pe</strong>. It is therefore important to apply correction factors for roughness<br />

when using this method.<br />

5-11 PRACTICAL CASES OF<br />

NON-NEWTONIAN SLURRIES<br />

The equations presented in the previous sections of this chapter are fairly complex and are<br />

based on so many assumptions that the practical engineer may feel lost. Some real examples<br />

are needed to guide the designer of non-Newtonian <strong>systems</strong>.<br />

5-11-1 Bauxite Residue<br />

HOMOGENEOUS FLOWS OF NONSETTLING SLURRIES<br />

Want et al. (1982) reviewed the design of a bauxite residue pi<strong>pe</strong>line for Alcoa Australia.<br />

The plant disposed 4.75 Mtpy (million tonnes <strong>pe</strong>r year) of alumina. Tests conducted on<br />

samples confirmed that the rheology of the <strong>slurry</strong> at concentrations in excess of 45% by<br />

weight could be expressed by the Carson equation:<br />

1/2 1/2 � w = � 0 + ���� �� 1/2 du<br />

�<br />

dr<br />

where du/dr is expressed by Equation 5-15 and � Equation 5-16.<br />

5.35


5.36 CHAPTER FIVE<br />

In the case of Kwina red mud, tests indicated that the modified consistency coefficient<br />

K� (Equation 5-18) was a complex function of the weight concentration:<br />

K� = 3.88 × 10 –18 (100 Cw) 10.57 Pa/s� and that the value of � was<br />

� = 2.55 × 106 (Cw × 100) –4.19 for Cw � 51%<br />

� = 6.685 (Cw × 100) –0.926 for Cw > 51%<br />

Applying the Dodge–Metzner model Want et al. (1982) applied Equation 5-20 to express<br />

the consumed power under turbulent conditions as:<br />

eT = 1.44 × 10 –4 150 – 100 Cw 2<br />

fn ��� (100 Cw) � 1.5<br />

(�Q) 3<br />

�5 D i<br />

in kW/km, and in laminar conditions as:<br />

eL = 0.393 K�� � 1+�<br />

� � 1+3�<br />

64�mQ(150 – 100Cw) 1<br />

��� � in kW/km<br />

3 × 10 Di<br />

Want et al (1982) discussed the thixotropic nature of red mud at high concentration, and<br />

the importance of flocculants and dis<strong>pe</strong>rsants on the rheology of this <strong>slurry</strong>. Referring to<br />

Figure 5-13, it is clear that there is a change in the pressure drop <strong>pe</strong>r unit length as the<br />

weight concentration is increased and the flow changes from turbulent to laminar. This<br />

drop in pressure to a minimum at such a transition is often misunderstood, particularly because<br />

it goes against the concepts examined in Chapter 4. The important parameters include<br />

the diameter of the pi<strong>pe</strong>, so that there is an optimum diameter, and an optimum<br />

weight concentration for a given tonnage of fine solids to be transported.<br />

Slurries may therefore be pum<strong>pe</strong>d at very high concentrations (Figure 5-14) using positive<br />

displacement pumps (Figure 5-15) over long distances, provided that the correct<br />

weight concentration is used near the turbulent to laminar transition region.<br />

4� Cw Example 5-11<br />

Using the data obtained by Want, examine pumping red mud bauxite residues at a s<strong>pe</strong>ed<br />

of 1.74 m/s in a 141 mm I.D. pi<strong>pe</strong> at weight concentrations of 45% and 60%. Determine<br />

the required power for a horizontal pi<strong>pe</strong>line, 3 km long. Assume a density of 1350 at 45%<br />

and 1800 at 60%.<br />

Solution<br />

At Cw = 45%:<br />

K� = 3.88 × 10 –18 (100Cw) 10.57<br />

K� = 1.157<br />

� = 2.55 × 106 (100 Cw) –4.19<br />

� = 0.302<br />

Q = AV = � × 0.25 × 0.1412 × 1.74 = 0.0272 m3 /s<br />

eT = 0.393 × 1.157� � 1.3<br />

� � 1.906<br />

64 × 1350 × 0.0272(150 – 45) 1<br />

���� �<br />

3 × 10 0.141<br />

4� × 0.45<br />

eT = 188 kW/km<br />

For 3 km, this is equivalent to 565 kw or 758 hp.


At C w = 60%:<br />

For 3 km, e L = 3123.9 kW or 4186 hp.<br />

HOMOGENEOUS FLOWS OF NONSETTLING SLURRIES<br />

K� = 3.88 × 10 –18 (60) 10.57 = 24.2<br />

� = 6.685(60) –0.926 = 0.151<br />

e T = 1041.3 kW/km<br />

� � 1.906 1<br />

eL = 0.393 × 24.2� � 1.3<br />

64 × 1800 × 0.0272 (150 – 60)<br />

���� �<br />

3 × 10 0.141<br />

4� × 0.6<br />

5.37<br />

FIGURE 5-13 Pressure drop <strong>pe</strong>r unit length for red mud tailings. (From F. M. Want et al.,<br />

1982. Reproduced by <strong>pe</strong>rmission of BHR Group.)


5.38 CHAPTER FIVE<br />

FIGURE 5-14 Slurry can be pum<strong>pe</strong>d in a non-Newtonian regime at high volumetric concentrations.<br />

(Courtesy of Geho Pumps.)<br />

5-11-2 Kaolin Slurries<br />

Slatter et al. (1996) reported that Kemblowski and Kolodziejski (1973) found that the<br />

Dodge and Metzner model did not well represent the flow of kaolin slurries. They derived<br />

the following empirical equation:<br />

and more generally:<br />

4f n =<br />

0.3164<br />

�0.25 ReMR<br />

4fn = �1/ReMR E<br />

�m ReMR


where E, m, and � are empirical parameters and functions of the apparent flow behavior<br />

index � (defined in Equation 5-16) and the modified Reynolds Number Re MR is defined<br />

by Equation 5-17.<br />

5-12 DRAG REDUCTION<br />

HOMOGENEOUS FLOWS OF NONSETTLING SLURRIES<br />

5.39<br />

FIGURE 5-15 The pumping of high concentration slurries and pastes may require positive<br />

displacement pumps. (Courtesy of Geho Pumps.)<br />

Ippolito and Sabatino (1984) showed that the addition of 3% salt to water tended to reduce<br />

the friction factor of bentonite sus<strong>pe</strong>nsions. Sauermann (1982) indicated that the addition<br />

of 0.2 kg/ton of tripolyphsophate (Na 5P 3O 10) to gold slimes at a weight concentration<br />

of 67.7% , and with particles smaller than 50 �m, reduced the pressure gradient in<br />

laminar flow by as much as 30%. Some viscosity reducing agents were discussed in


5.40 CHAPTER FIVE<br />

Chapter 3.There are, however, very little data published on other methods for reducing<br />

friction losses of non-Newtonian slurries.<br />

Schowalter (1977) discussed some as<strong>pe</strong>cts of drag reduction in non-Newtonian slurries<br />

and reported certain cases of mixtures with a pressure drop actually lower than that of<br />

water.<br />

5-13 PULP AND PAPER<br />

Pulp and pa<strong>pe</strong>r pump slurries behave as non-Newtonian slurries. The following equations<br />

have been reported by the Cameron Hydraulic Data book of IDP (1995), based on work<br />

at the University of Maine. A modified Reynolds number is defined as<br />

D I 0.205 Vs�g<br />

ReMR = (5-81)<br />

where<br />

C = % consistency of the pulp, oven dry<br />

g = 32.2 ft/s<br />

� = density in slugs/ft3 �� 1.157 C<br />

Vs = velocity, defined in ft/sec<br />

A modified friction factor is defined as<br />

3.97<br />

f = � (5-82)<br />

1.636<br />

ReMR<br />

A s<strong>pe</strong>cial equation for friction losses is therefore defined as<br />

fV 2 LK 0<br />

Hf = � (5-83)<br />

DI<br />

The correction factor K0 de<strong>pe</strong>nds on the ty<strong>pe</strong> of pulp. It is considered to be 1.00 for unbleached<br />

sulfite softwood, 0.90 for bleached sulfite softwood, 0.90 for unbleached kraft<br />

softwood, 0.90 for soda hardwood, 0.90 for reclaimed fiber, 1.0 for presteamed groundwood–softwood,<br />

and 1.42 for stoned groundwood–softwood. (IDP 1995).<br />

The following program calculates friction factors.<br />

10 PRINT “program for stock , pulp and pa<strong>pe</strong>r”<br />

PRINT “based on the curves of the University of Maine”<br />

PRINT “ which are correlations to the data of Brecht and Heller”<br />

pi = 4 * ATN(1)<br />

INPUT “name of project”; pr$<br />

INPUT “name of client “; nc$<br />

INPUT “date of calculations”; dat$<br />

INPUT “ ty<strong>pe</strong> of pulp”; pul$<br />

INPUT “consistency of pulp in <strong>pe</strong>rcent”; CS<br />

C = CS/100<br />

15 INPUT “ choose between (1) SI units (m) and (2) US units (feet) “; NB<br />

IF ABS(NB) < 1 THEN GOTO 15<br />

IF ABS(NB) > 2 THEN GOTO 15<br />

IF (NB > 1) AND (NB < 2) THEN GOTO 15<br />

GOSUB conversion<br />

18 IF NB = 1 THEN INPUT “inner diameter (m)”, D1M<br />

IF NB = 2 THEN INPUT “inner diameter (ft)”; D1US


IF NB = 1 THEN d1 = D1M * dl<br />

IF NB = 1 THEN D1US = D1M/.3048<br />

IF NB = 2 THEN d1 = D1US * dl<br />

IF NB = 1 THEN INPUT “pulp flow rate (m3/hr)”, qm<br />

IF NB = 2 THEN INPUT “pulp flow rate in USgpm”; qus<br />

IF NB = 1 THEN q = qm/60000<br />

IF NB = 2 THEN q = qus * 3.785/60000<br />

a = .25 * pi * d1 ^ 2<br />

v = q/a<br />

PRINT USING “pulp s<strong>pe</strong>ed = ###.### m/s”; v<br />

vus = v/.3048<br />

PRINT USING “PULP SPEED = ##.### ft/s”; vus<br />

IF (C > .03) AND (vus > 8) THEN PRINT “s<strong>pe</strong>ed exceeds 8 ft/s please use<br />

larger pi<strong>pe</strong> size”<br />

IF C > .03 THEN GOTO 25<br />

IF (C < .02) AND (vus > 10) THEN PRINT “s<strong>pe</strong>ed exceeds 10 ft/s please use<br />

larger”<br />

IF C < .02 THEN GOTO 25<br />

IF vus > 9 THEN PRINT “SPEED EXCEEDS 9 FT/S, PLEASE USE LARGER PIPE”<br />

25 INPUT “DO YOU WANT TO USE A LARGER PIPE (Y/N)”; p$<br />

IF p$ = “Y” OR p$ = “y” THEN GOTO 18<br />

IF NB = 1 THEN INPUT “DENSITY OF STOCK IN KG/M3 (OFTEN ASSUMED TO BE 1000<br />

KG/M3 “; DENSM<br />

IF NB = 1 THEN DENSUS = DENSM * (62.4/1000)<br />

IF NB = 2 THEN INPUT “DENSITY OF STOCK IN LBS/CU.FT (OFTEN ASSUMED TO BE<br />

62.4 LBS/CU.FT”; DENSUS<br />

REM0 = D1US ^ .205 * vus * DENSUS/CS ^ 1.157<br />

FM0 = 3.97/REM0 ^ 1.636<br />

PRINT USING “MODIFIED REYNOLDS NUMBER = #######”; REM0<br />

PRINT USING “MODIFIED FRICTION FACTOR = #.####”; FM0<br />

PRINT “ please choose between the following pulps”<br />

PRINT “ 1- unbleached sulfite “<br />

END<br />

conversion:<br />

IF NB = 1 THEN dl = 1<br />

IF NB = 2 THEN dl = .3048<br />

IF NB = 1 THEN ql = 3875/60<br />

IF NB = 2 THEN ql = 60/3875<br />

RETURN<br />

5-14 CONCLUSION<br />

HOMOGENEOUS FLOWS OF NONSETTLING SLURRIES<br />

5.41<br />

The world of non-Newtonian flows is very complex and encompasses very different<br />

flows, including pulp and pa<strong>pe</strong>r, food, plastics, and clays. The use of various equations<br />

develo<strong>pe</strong>d by researchers do not yield the same values of the friction coefficient. The<br />

problem is compounded by the fact that many rheological tests are conducted in tubes and<br />

in laminar flows, yielding values of consistency factor K and exponent n outside the range<br />

of turbulent flows. The practical engineer is often left to use his engineering judgment to<br />

use the appropriate equation. Pro<strong>pe</strong>r tests of the <strong>slurry</strong> flow at the correct range of shear<br />

stresses are essential to avoid errors. Various reference books on the subject have equations<br />

similar to Equations 5-20 to 5-25 without emphasizing the limitations to their use.<br />

Because many of the equations for friction loss factors are implicit and require iteration<br />

methods, the use of <strong>pe</strong>rsonal computers is encouraged.<br />

The flow of non-Newtonian slurries is complex and requires a significant energy in-


5.42 CHAPTER FIVE<br />

put. Methods have been develo<strong>pe</strong>d over the years to reduce friction losses. These include<br />

dilution, reduction of volumetric concentration of solids, removal of all flocculants, provisions<br />

for air purging of the pi<strong>pe</strong>line, addition of high-as<strong>pe</strong>ct-ratio fibers, addition of deflocculants<br />

(soluble ionic compounds), reduction of the angularity of particles, and addition<br />

of viscosity reducing agents. Not all these methods are always possible, and some<br />

require capital investment.<br />

It is ho<strong>pe</strong>d that the various worked examples in this book will help the practical engineer<br />

to design <strong>slurry</strong> pi<strong>pe</strong>lines. It may be necessary to use more than one method and<br />

compare results. There are many advantages to pumping slurries at high concentrations,<br />

such as concentrates from process plants, food pastes, and ceramic <strong>slurry</strong> for the manufacture<br />

of new materials.<br />

5-15 NOMENCLATURE<br />

a Nondimensional parameter and function of Hedstrom number<br />

A Factor for friction in the Churchill equation<br />

Ar Archimedean number<br />

b Nondimensional parameter<br />

B Factor for friction in the Churchill equation<br />

C1 Nondimensional power law parameter<br />

CD Drag coefficient<br />

Cv Volume fraction of solid particles in the <strong>slurry</strong> mixture<br />

Cw Weight concentration<br />

d85 Particle diameter passing 85% (m)<br />

dp Diameter of particle<br />

dP/dz Pressure gradient <strong>pe</strong>r unit length<br />

dx Equivalent roughness<br />

Di Conduit inner diameter (m)<br />

eT Consumed energy<br />

E Empirical coefficient<br />

fD Darcy friction factor<br />

fL Laminar component of fanning friction factor for a Bingham plastic<br />

fN Fanning friction factor<br />

fNC Fanning friction at transition between laminar and turbulent flow<br />

fNL Laminar component of fanning friction factor<br />

fNLY Laminar component of fanning friction factor for a yield pseudoplastic<br />

fNPL Fanning friction factor for a pseudoplastic in a laminar regime<br />

Tomita laminar friction factor<br />

fPLT fT fTR Turbulent component of fanning friction factor<br />

Fr<br />

Darby friction factor for transition from laminar to turbulent flows with power<br />

law fluids<br />

Froude number<br />

g Acceleration due to gravity (9.8 m/s2 )<br />

gc Conversion from slugs to pounds mass in USCS units<br />

He Hedstrom number for Bingham plastic<br />

He mod<br />

Modified Hedstrom for yield pseudoplastic<br />

j Herschel–Bulkley parameter<br />

K Coefficient of consistency for power law fluids<br />

K� Modified coefficient of consistency for power law fluids


L Length of conduit or pi<strong>pe</strong><br />

m Exponent in Darby’s equation for fanning friction factor calculations<br />

n Flow behavior index for pseudoplastic flows<br />

P Pressure<br />

Plasticity number<br />

PL Pst Start-up pressure to pump a non-Newtonian <strong>slurry</strong><br />

Q Flow rate<br />

R� Chilton and Stainsby proposed modified Reynolds number<br />

Re Reynolds number<br />

ReB Reynolds number for a Bingham plastic using the coefficient of rigidity for viscosity<br />

ReBc Critical transition Reynolds number for a Bingham plastic using the coefficient<br />

of rigidity for viscosity<br />

Rec Reynolds number at transition<br />

Remod Modified Hedstrom number for yield pseudoplastic<br />

ReMR Modified Reynolds number for a power law fluid<br />

RePL Tomita Reynolds number for a power law fluid<br />

RePLC Tomita Reynolds number for a power law fluid at transition<br />

Rer Slatter Reynolds number<br />

Rep Particle Reynolds number<br />

Uf Friction velocity<br />

V S<strong>pe</strong>ed<br />

VN Newtonian velocity<br />

Vr Slip velocity<br />

Vs Velocity of solids<br />

Vt Terminal velocity of falling particles<br />

VTR Transition velocity from laminar to turbulent flows<br />

Ratio of slip velocity to <strong>slurry</strong> s<strong>pe</strong>ed<br />

W r<br />

x Ratio of the yield stress to the wall shear stress<br />

x c<br />

Ratio of the yield stress to the wall shear stress at the transition from laminar to<br />

turbulent flow<br />

Z Settling factor for a non-Newtonian fluid<br />

Subscripts<br />

L Liquid<br />

m Mixture<br />

p Particle<br />

Greek symbols<br />

� Function for use of laminar and friction factors<br />

� Increment<br />

� Concentration by volume in decimal points<br />

�P Pressure drop<br />

��m Density change for the mixture<br />

� Density<br />

�L Density of liquid carrier in kg/m3 �m Density of <strong>slurry</strong> mixture in Kg/m3 � s<br />

Density of solids<br />

HOMOGENEOUS FLOWS OF NONSETTLING SLURRIES<br />

� Modified flow behavior index for pseudoplastic flows<br />

� Shear strain<br />

d�/dt Wall shear rate or rate of shear strain with res<strong>pe</strong>ct to time<br />

5.43


5.44 CHAPTER FIVE<br />

� Linear roughness (m)<br />

� Coefficient of rigidity of a non-Newtonian fluid, also called Bingham plastic viscosity<br />

�� Coefficient of rigidity at high shear rate<br />

� Carrier liquid dynamic viscosity<br />

�a Apparent viscosity of a pseudoplastic fluid<br />

�e Effective pi<strong>pe</strong>line viscosity<br />

�e�� Wilson–Thomas effective viscosity<br />

�L Viscosity of liquid carrier<br />

�p Effective pi<strong>pe</strong>line viscosity for pseudoplastic<br />

� * Effective viscosity for Hagen-Poiseuille fluid<br />

� �<br />

Bingham plastic limiting viscosity of <strong>slurry</strong> mixtures (Poise)<br />

� Empirical function of the power exponent n<br />

� Pythagoras number (ratio of circumference of a circle to its diameter)<br />

� Shear stress at a height y or at a radius r<br />

�0 Yield stress for a Bingham plastic<br />

�w Wall shear stress<br />

�yp Yield stress for pseudoplastic<br />

5-16 REFERENCES<br />

Abulnaga, B. E. 1997. Slurcal—Computer Program for Non-Newtonian Flows. Fluor Daniel Wright<br />

Engineers, Vancouver, BC, Canada (unpublished).<br />

Aral, B. K., and D. M. Kaylon. 1994. Effects of tem<strong>pe</strong>rature and surface roughness on time de<strong>pe</strong>ndent<br />

development of wall slip in steady torsional flow of concentrated Sus<strong>pe</strong>nsions. Journal of<br />

Rheology, 38, 957–972.<br />

Bowen, R. L. 1961. Chemical Engineering, 143–150.<br />

Buckingham, E. 1921. On plastic flow through capillary tubes. ASTM Proceedings, 21, 1154.<br />

Chilton, R. A., and R. Stainsby. 1998. Pressure loss equations for laminar and turbulent non-<br />

Newtonian pi<strong>pe</strong> flow. Journal of Hydraulic Engineering, 124, 5, 522–529.<br />

Clifton, R. A, and R. Stainsby. 1998. Pressure loss for laminar and turbulent non-Newtonian pi<strong>pe</strong><br />

flow. Journal of Hydraulic Engineering, 124, 5, 522–529.<br />

Churchill, S. W. 1977. Friction factor equation spans all fluid flow regimes. Chemical Engineering<br />

84, 7, 91–92.<br />

Darby, R. 2000. Pressure drop of non-Newtonian slurries, a wider path. Chemical Engineering, 107,<br />

5, 64–67.<br />

Darby, R. 1981. How to predict the friction factor for flow of Bingham plastics. Chemical Engineering,<br />

88, 26, 59–61.<br />

Darby, R., R. Mun, and V. Boger. 1992. Prediction friction loss in <strong>slurry</strong> pi<strong>pe</strong>s. Chemical Engineering,<br />

September.<br />

Dodge, D. W., and A. B. Metzner. 1959. Turbulent flow of non-Newtonian <strong>systems</strong>. Am. Inst. Chem.<br />

Engr., 5, 2, 189–204.<br />

Edwards, M. F., M. S. M. Jadallah, and R. Smith. 1985. Head losses in pi<strong>pe</strong> fittings at Low Reynolds<br />

Number. Chem. Engr, Res. Des., 63, 1, 43–50.<br />

Govier, G. W., and K. Aziz. 1972. The Flow of Complex Mixtures in Pi<strong>pe</strong>s. New York: Van Nostrand<br />

Reinhold.<br />

Hanks, R. W. 1962. A Generalized Criterion for Laminar–Turbulent Transition in the Flow of Fluids.<br />

Union Carbide report.<br />

Hanks, R. W., and D. R. Pratt. 1967. On the flow of Bingham plastic slurries in pi<strong>pe</strong>s and between<br />

parallel plates. Soc. Petr. Eng. Journal, 7, 342–346.<br />

Hanks, R. W., and B. H. Dadia. 1971. Theoretical analysis of the turbulent flow of non-Newtonian<br />

slurries in pi<strong>pe</strong>s. American Journal of Chemical Engineering, 17, 554–557.


HOMOGENEOUS FLOWS OF NONSETTLING SLURRIES<br />

5.45<br />

Hanks, R. W., and B. L. Ricks. 1975. Transitional and turbulent pi<strong>pe</strong>flow of pseudoplastic fluids.<br />

Journal of Hydronautics, 9, 39–44.<br />

Hedstrom, B. O. A. 1952. Flow of plastic materials in pi<strong>pe</strong>s. Ind. Eng. Chem., 44, 651–656.<br />

Herschel, W. H., and R. Bulkley. 1928. Measurements of consistency as applied to rubber benzene<br />

solution. Proc. ASTM, 26, Part 2, 621–633.<br />

Herzog, R. L., and K. Weissenberg. 1928. Kolloid Z, 46, 277.<br />

Heywood, N. I. 1991. Pi<strong>pe</strong>line design for non-settling slurries. In Slurry Handling, Brown, N. P., and<br />

N. I. Heywood (Eds.). New York: Elsevier Applied Sciences.<br />

Heywood, N. I., D. C. H. Cheng, and A. J. Carlton. 1992. Slurry <strong>systems</strong>. In Piping Design Handbook,<br />

McKetta, J. J. (Ed.). New York: Marcel Dekker, pp. 585–622<br />

Heywood, N. I., and J. F. Richardson. 1978. Rheological behavior of flocculated and dis<strong>pe</strong>rsed<br />

kaolin sus<strong>pe</strong>nsions in pi<strong>pe</strong> flow. Journal of Rheology, 22, 6, 559–613.<br />

IDP (now called Flowserve). 1995. Cameron Hydraulic Data. NJ: IDP.<br />

Ippolito, M., and C. Sabatino. 1984. Rheological behavior and friction resistance of colloidal aqueous<br />

sus<strong>pe</strong>nsions. In Proceedings of the IXth International Congress on Rheology, Mexico.<br />

Mexico: Universidad Nacional Autonoma de Mexico.<br />

Irvine. 1988. Ex<strong>pe</strong>rimental measurements of isobaric thermal expansion coefficients of Non-Newtonian<br />

fluids. Heat Transfer, 1, 2, 155–163.<br />

Johnson, M. 1982. Non-Newtonian fluid system design. Some problems and their solutions. In 8th<br />

International Conference on the Hydraulic Transport of Solids in Pi<strong>pe</strong>. Johanesburg, South<br />

Africa. Cranfield, UK: BHR Group.<br />

Kemblowski, Z., and J. Kolodziejski. 1973. Flow resistances of non-Newtonian fluids in transitional<br />

and turbulent flow. Int. Chem. Eng., 13, 265–279.<br />

Kenchington, J. M. 1972. In Proceedings of the 2nd International Conference on Hydraulic Transportation<br />

of Solids.Cranfield, UK: BHR Group.<br />

Metzner, A. B., and J. C. Reed. 1955. Flow of non-Newtonian laminar, transition and turbulent regions.<br />

Am. Inst. Chem. Eng. Journal, 1, 4, 434.<br />

Molerus, O. 1993. Principles of Flow in Dis<strong>pe</strong>rse Systems. London: Chapman and Hall.<br />

Mooney, M. J. 1931. Explicit formulas for slip and fluidity. Journal of Rheology, 2, 2, 210–222.<br />

Nunez, G. A., M. Briceno, C. Mata, and H. Rivas. 1996. Flow characteristics of concentrated emulsions<br />

of very viscous oil in water. The Journal of Rheology, 40, 3, 405–423.<br />

Park, J. T., R. J. Munnheimer, T. A. Grimley, and T. B. Morrow. 1989. Pi<strong>pe</strong> flow measurements of a<br />

transparent Non-Newtonian <strong>slurry</strong>. Journal of Fluids Engineering, 111, 331–336.<br />

Porkryvailo, N. A., and Y. G. Grozberg. 1995. Investigation of structure of turbulent wall flow of<br />

clay sus<strong>pe</strong>nsions in channel with electro diffusion method. In Proceedings of the 8th International<br />

Conference on Transport and Sedimentation of Solid Particles, Prague, Czech Republic.<br />

Rabinowitsch, B. 1929. Veber die viskositat und elastizitat von solen. Z. Phisik Chem. Ser. A., 145,<br />

1–26.<br />

Ryan, N. M., and M. M. Johnson. 1959. Transition from laminar to turbulent flows in pi<strong>pe</strong>s. Amer.<br />

Inst. of Chem. Engr., 5, 433–435<br />

Sauermann, H. B. 1982. The Influence of particle diameter on the pressure gradients of gold slimes<br />

pumping In Proceedings of the 8th International Conference on the Hydraulic Transport of<br />

Solids in Pi<strong>pe</strong>s. Jahannesburg, South Africa, August 1982, Pa<strong>pe</strong>r E1, pp. 241–246. Cranfield,<br />

UK: BRHA Group.<br />

Schowalter, W. R. 1977. Mechanics of Non-Newtonian Fluids. New York: Pergamon Press.<br />

Slatter, P. T., G. S. Thorvaldsen, and F. W. Petersen. 1996. Particle roughness turbulence. In Proceedings<br />

of the 13th International Hydrotransport Symposium on Slurry Handling and Pi<strong>pe</strong>line<br />

Transport, Johannesburg, South Africa. Cranfield, UK: BRHA Group.<br />

SRC. 2000. Slurry Pi<strong>pe</strong>line Course, May 15–16, 2000, Saskatchewan Research Centre, Saskatoon,<br />

Canada.<br />

Szilas, A. P., E. Bobok, and L. Navratil. 1981. Determination of turbulent pressure loss Non-Newtonian<br />

oil flow in rough tubes. Rheol Acta, 20, 487–496.<br />

Thomas, A. D., and K. C. Wilson. 1987. New analysis of non-Newtonian turbulent flow, yield power<br />

law fluids. Canadian Journal of Chemical Engineering, 65, 335–338.<br />

Tomita, Y. 1959. On the fundamental formula of non-Newtonian flow. Bulletin of the Japan. Soc.<br />

Mech.. Engr., 2, 7, 469–474.


5.46 CHAPTER FIVE<br />

Torrance, B. McK. 1963. Friction factors for turbulent non-Newtonian fluid flow in circular pi<strong>pe</strong>s.<br />

South African Mechanical Engineer, 13, 4, 89–91.<br />

Want, F. M., P. M. Colombera, Q. D. Nguyen, and D. V. Boger. 1982. Pi<strong>pe</strong>line design for the transport<br />

of high-density bauxite residue slurries. In Proceedings of the 8th International Conference<br />

on the Hydraulic Transport of Solids in Pi<strong>pe</strong>s, Johannesburg, South Africa. Cranfield,<br />

UK: BRHA Group.<br />

Wasp, E., J. Penny, and R. Ghandi. 1977. Solid-liquid flow <strong>slurry</strong> pi<strong>pe</strong>line transportation. Clausthal,<br />

Germany: Trans Tech Publications.<br />

Wilson, K. C. 1985. A new analysis of turbulent flow of non-Newtonian fluids. Canadian Journal of<br />

Chemical Engineering, 63, 539–546.<br />

Further readings<br />

Abulnaga, B. E. 1997. Channel 1.0 Computer Program For an O<strong>pe</strong>n Channel Slurry Flow. Fluor<br />

Daniel Wright Engineers. Vancouver, BC, Canada. Internal report.<br />

Al Fariss, T. and K. L. Pinder. 1987. Flow through porous media of a shear-thinning liquid with yield<br />

stress. Canadian Journal of Chemical Engineering, 65, 391–405.<br />

Bouzaiene, R., and D. Hassani-Ferri. 1992. A selection of pressure loss predictions based on <strong>slurry</strong>/backfill<br />

characterization and flow conditions. C.I.M. Bulletin, 85, 959, 63–68.<br />

Chlabra, R. P., J. F. Richardson, and R. Darby. 2000. Non-Newtonian flow in process industry, fundamentals<br />

and engineering application. Chemical Engineering, 107, 4,<br />

Draad, A. A., G. D. C. Kuiken, and F. T. M. Nieuwstadt. 1998. Laminar turbulent transition in pi<strong>pe</strong><br />

flow for Newtonian and non-Newtonian fluids. Journal of Fluid Mechanics, 377, D25,<br />

267–312.<br />

Hedstrom, B. O. A. 1952. Flow of plastic materials in pi<strong>pe</strong>s. Journal of Industrial Engineering<br />

Chemistry, 44, 651–656.<br />

Sandall, O. C., O. T. Hanna, and K. Amurath. 1986. Ex<strong>pe</strong>riments on turbulent non-Newtonian mass<br />

transfer in a circular tube. Am. Inst. Chem. Eng. Journal, 32, 2095–2098.<br />

Steffe, J. F., and R. G. Morgan. 1986. Pi<strong>pe</strong>line design and pump selection for non-Newtonian fluid<br />

foods. Food Technology, 40, 78–85.<br />

Wheeler, J. A., and E. H. Wissler. 1965. Friction factor: Reynolds number relation for the steady<br />

flow of pseudoplastic fluids through rectangular ducts. Am. Inst. Chem. Eng. Journal, 11,<br />

207–216.


PART TWO<br />

EQUIPMENT AND<br />

PIPELINES


CHAPTER 7<br />

COMPONENTS OF<br />

SLURRY PLANTS<br />

7-0 INTRODUCTION<br />

In Chapter 1, a typical circuit of a mineral process plant was presented. In Chapters 3<br />

through 6, the theory of <strong>slurry</strong> flows was examined in detail for different rheology and<br />

regimes. To achieve such complex flows, a number of important pieces of machinery,<br />

such as mills, pumps, and valves, and drop boxes are needed. Together they form the <strong>slurry</strong><br />

preparation plant at the start of the pi<strong>pe</strong>line and sometimes the <strong>slurry</strong> dewatering plant<br />

when the concentrate or solids must be dried out for shipping, smelting, or burning as a<br />

fuel. Their design is often complex and must account for wear and <strong>pe</strong>rformance.<br />

In simple layman’s terms, rocks that contain ores may be delivered in fairly large<br />

pieces. These rocks may be obtained by blasting, s<strong>pe</strong>cial hydraulic jack hammers, excavators,<br />

etc. (Figure 7-1). These large rocks need to be reduced to sufficiently small particles<br />

to extract the ores—from as large as a few hundred millimeters (or dozens of<br />

inches) down to a few millimeters or fractions of inches. This is done by a number of<br />

steps, such as crushing, milling, grinding, screening, cycloning, vibrating, etc. Milled<br />

rocks are then transported in <strong>slurry</strong> form and treated in different circuits such as flotation,<br />

acid or cyanide leaching, and classification circuits. The concentrate may then be<br />

thicked further for transportation to its final destination. The tailings are disposed of in<br />

dedicated ponds.<br />

The design of mineral processing plants has been the subject of numerous books, and<br />

s<strong>pe</strong>cialized books have been written for each piece of equipment. In this chapter, some of<br />

the most important components of <strong>slurry</strong> <strong>systems</strong> will be introduced, with sufficient information<br />

for the <strong>slurry</strong> engineer to appreciate the discharge from each ty<strong>pe</strong> of equipment.<br />

The next two chapters are devoted to pumps and valves and Chapter 10 is devoted to materials<br />

for manufacturing. It would be beyond the sco<strong>pe</strong> of this book to dwell on the<br />

chemistry of each process.<br />

7-1 ROCK CRUSHING<br />

Rock crushing is not part of the <strong>slurry</strong> circuit but is more of a preparatory step to the formation<br />

of slurries. Crushing will therefore be reviewed briefly, as it is outside the sco<strong>pe</strong><br />

of this <strong>handbook</strong>.<br />

7.3


7.4 CHAPTER SEVEN<br />

FIGURE 7-1 Excavation is a primary source of materials for a mineral processing plant.<br />

[Courtesy of Metso Minerals (formerly known as the companies Nordberg and Svedala).]<br />

Solid comminution is the process of reducing the size of particles. Two comminution<br />

ty<strong>pe</strong>s are considered:<br />

1. Dry comminution generally reduces rocks down to a diameter of 25 mm (1 in), by impact<br />

and mechanical compression. This process involves jaw crushing, gyratory crushing,<br />

cone crushing, and grinding using rod mills and ball mills.<br />

2. Wet comminution generally reduces 25 mm (1 in) particles down to very fine sizes by<br />

grinding and attrition in <strong>slurry</strong> form. This process involves semiautogenous mills, autogenous<br />

mills, ball mills, hydrocyclones, columns, etc.<br />

Comminution via a machine is measured by the reduction ratio, defined as 80% of the<br />

particle size at the feed (F e80) to 80% of the particle size at the output (C r80).<br />

The feed to a grinding mill must be crushed to a size appropriate to the grinding<br />

process. Semiautogenous mills require little crushing; ball mills require a finer crushing.<br />

A method of ore preparation that is now limited to narrow ore seams or veins in underground<br />

mines is the so-called “run of the mine milling.” It consists of blasting the rocks<br />

into lumps, usually of the order if 300 mm (12 inch) or larger. The most common approach,<br />

however, is to crush the mined rock to an acceptable size.<br />

7-1-1 Primary Crushers<br />

Primary crushers absorb any size rocks (de<strong>pe</strong>nding on the o<strong>pe</strong>ning at the inlet) and reduce<br />

their size down to 50–150 mm (2–6 in). Primary crushers are classified as:


� Jaw crushers<br />

� Gyratory crushers<br />

� Impact crushers<br />

COMPONENTS OF SLURRY PLANTS<br />

Some mines try to reduce the cost of crushing by blasting the rocks from mountains<br />

and hills.<br />

Crushing is essentially a process of reducing the size of a stone down to 25 mm (1 in)<br />

(Figure 7-2). As this is difficult to achieve in a single stage, it is often encompassed in two<br />

or three steps. The stones go through a cycle of primary crushing, secondary crushing,<br />

and tertiary crushing. S<strong>pe</strong>cial machines have been develo<strong>pe</strong>d for each step of crushing<br />

(Figure 7-3).<br />

7-1-1-1 Jaw Crushers<br />

These machines o<strong>pe</strong>rate by compressing the rocks between a fixed plate and a moving<br />

jaw (Figure 7-4). The rocks are fed from the top of the crusher. The fixed jaw or plate is<br />

usually attached to the wall of a cavity. Through an eccentric mechanism or crankshaft, a<br />

moving jaw presses the rocks against the walls of the crusher. Generally, the following<br />

two ty<strong>pe</strong>s of machines are used:<br />

1. In the overhead eccentric jaw crusher, also known as the single toggle crusher, the<br />

moving plate is forced against the stationary plate by an eccentric mechanism driving<br />

at its top, as well as by the rocking of a toggle connected to the bottom of the<br />

moving plate.<br />

FIGURE 7-2 Crushing is an essential step in handling hard rock, gravel, and mining ores as<br />

well as for recycling. [Courtesy of Metso Minerals (formerly known as the companies Nordberg<br />

and Svedala).]<br />

7.5


7.6 CHAPTER SEVEN<br />

fixed jaw<br />

out<br />

feed<br />

(a) Jaw crusher<br />

feed<br />

bowl Head or mantle<br />

(c) Impact crusher<br />

Pivoting jaw<br />

pitman<br />

bowl<br />

inclined bowl<br />

FIGURE 7-3 Principles of crushing.<br />

feed<br />

(b) Gyratory crusher<br />

feed<br />

(b) Cone crusher<br />

Head or mantle<br />

FIGURE 7-4 Cross-sectional representation of a jaw crusher. [Courtesy of Metso Minerals<br />

(formerly known as the companies Nordberg and Svedala).]<br />

cone


COMPONENTS OF SLURRY PLANTS<br />

2. The blake jaw crusher features a moving plate that pivots at the top but is oscillated at<br />

the bottom.<br />

The dimensions and sha<strong>pe</strong> of the plates affect the <strong>pe</strong>rformance of the crusher. The<br />

smaller the discharge gap, or required output size, the lower the tonnage from the crusher.<br />

Jaw crushers work best on rocks that are not flat or slabs. With a feed o<strong>pe</strong>ning of 1.67 ×<br />

2.13 m (66 × 84 in) and a discharge gap of 200 mm (8 in), the crusher can handle a capacity<br />

of 800 tph.<br />

The walls and moving blade of the crusher are lined with a hard metal such as manganese<br />

steel. The liners are removable for repairs once worn out. The liners may be flat,<br />

plain, or ribbed.<br />

The final output size of crushed particles de<strong>pe</strong>nd on the setting of the plates (Figure 7-<br />

5). Curves shown in Figure 7-5 indicate, for example, that for a closed setting of 100 mm<br />

(4 in) the size particles will be at a maximum of 160 mm (6.375 in) with a significant portion<br />

of particles smaller than 50 mm (2 in).<br />

7-1-1-2 Gyratory Crushers<br />

These machines o<strong>pe</strong>rate on the principle of compressing the rocks in a cone (Figure 7-6)<br />

The rocks fall into the cavity from the top. The moving part is an eccentric cone. The<br />

FIGURE 7-5 The size of the output from jaw crushers de<strong>pe</strong>nds on the plate setting. If the<br />

closed side setting (c.s.s) is 100 mm (4�), the maximum product size is 160 mm (6 3 – 8�) and the<br />

portion of fraction under 50 mm (2�) is approximately 35%. [Courtesy of Metso Minerals (formerly<br />

known as the companies Nordberg and Svedala).]<br />

7.7


7.8 CHAPTER SEVEN<br />

Mainshaft sleeve<br />

Spider bushing<br />

Spider arm guard<br />

Head nut<br />

Spider<br />

Concave fifth row<br />

Concave fourth row<br />

Concave third row<br />

Concave second row<br />

Concave first row<br />

Inner deflector ring<br />

Arm guard (inner)<br />

Arm guard (outer)<br />

Bottom shell<br />

Tie rod nut<br />

Gear housing shield<br />

Positive air pressure<br />

Eccentric<br />

Seal ring<br />

Eccentric support<br />

Hydraulic cylinder<br />

Cylinder sleeve<br />

Cylinder shield<br />

Piston cap<br />

Cylinder head<br />

Transmitter<br />

Spider cap<br />

Mainshaft<br />

Retainer bar<br />

Guide bushing<br />

Seal retainer<br />

Tie rod nut<br />

Top shell<br />

Up<strong>pe</strong>r mantle<br />

Tie rod<br />

Lower mantle<br />

Floating ring bracket<br />

Oil deflector ring<br />

Dust seal bonnet<br />

Floating ring<br />

Floating ring retainer<br />

Outer bushing<br />

Pinion<br />

Inner busing<br />

Countershaft box<br />

Countershaft<br />

Balanced gear<br />

Eccentric thrust washer<br />

Eccentric thrust bearing<br />

Swivel plate<br />

Socket plate<br />

Thrust plate<br />

FIGURE 7-6 Cross-sectional drawing of a primary gyratory crusher. (Courtesy of Sandvik.)<br />

rocks enter on the largest corner of the cavity but are compressed as the eccentric cone rotates.<br />

The outside cone is sometimes called the bowl, and the rotating cone is called the<br />

mantle. The bowl reduces in diameter toward the bottom, whereas the mantle increases in<br />

diameter with depth in the opposite direction.<br />

Gyratory crushers are preferred for slabs or flat-sha<strong>pe</strong>d rocks as they snap the rock<br />

better. Gyratory crushers are manufactured to handle tonnage flows up to 3500 tph. Sandvik<br />

purchased the line of Nordberg mobile primary gyratory crushers (Figure 7-7) that<br />

can be moved from one site to another as the mine expands.<br />

7-1-1-3 Impact Crushers<br />

These machines o<strong>pe</strong>rate on the principle of a set of rotating hammers hitting against the<br />

rocks. The hammers are fixed to a cylinder. The feed is from the top and as the rocks feed<br />

in, they fall between a breaker plate and the rotating cylinder. The hammers produce the<br />

required impact to chip the rocks. Impact crushers work best on rocks that are neither<br />

abrasive nor silica-rich, as these cause rapid wear of the hammers. Metso Minerals manufactures<br />

impact crushers (Figure 7-8) for primary and secondary crushing. Figure 7-9<br />

shows typical gradation curves.


7-2 SECONDARY AND<br />

TERTIARY CRUSHERS<br />

Crushing the rocks is often achieved in two or three stages. The secondary and tertiary<br />

crushing machines resemble the machines used during primary crushing. They consist of<br />

vertical cone crushers or horizontal cylinder crushers. The former ty<strong>pe</strong> is the most widespread.<br />

7-2-1 Cone Crushers<br />

COMPONENTS OF SLURRY PLANTS<br />

FIGURE 7-7 Large mobile gyratory crushers are designed with a s<strong>pe</strong>cial frame and wheels<br />

to <strong>pe</strong>rmit relocation from one area of the mine to another. (Courtesy of Sandvik.)<br />

Cone crushers o<strong>pe</strong>rate on the same principle as gyratory crushers. This allows a gradual<br />

reduction of the area between the two cones. The rotating cone or mantle is inclined, thus<br />

providing a combination of impact loads and compression loads. By comparison with the<br />

gyratory crusher, the outer bowl is inverted, and the mantle rotates at much higher s<strong>pe</strong>eds.<br />

There are two ty<strong>pe</strong>s of cone crushers:<br />

1. The standard ty<strong>pe</strong> (for secondary crushing)<br />

2. The short head ty<strong>pe</strong> (for tertiary crushing)<br />

7.9


7.10 CHAPTER SEVEN<br />

FIGURE 7-8 Cross-sectional cut through an impact crusher. (Courtesy of Sandvik.)<br />

The two ty<strong>pe</strong>s of cone crushers have different bowl sha<strong>pe</strong>s. The standard has a wider feed<br />

and is used for larger stones. The short head has a more shallow feed and tighter space<br />

surrounding the mantle. The short head is therefore used for finer crushing.<br />

Because of the continuous wear of the surfaces, adjustment of the cone crusher is essential.<br />

By measuring power on a continuous basis, a feedback loop readjusts the mantle.<br />

Screens on the output of the crusher facilitate the separation of coarse and fine stones. In a<br />

closed circuit, the coarser stones are returned to the crusher. The fine stones could clog<br />

the crusher and must be removed.<br />

The diameter of cone crushers may be as low as 0.91 m (36 in) for a capacity of 50–80<br />

tph, or as high 2.13 m (84 in) for a capacity of 500–1100 tph. The finer the output, the<br />

smaller is the tonnage.


Figure 7-10 presents a cross-sectional drawing of the Metso Minerals cone crusher<br />

and Figure 7-11 shows gradation curves of the output from HP cone crushers. Metso Minerals<br />

manufactures complete portable cone/screen plants (Figure 7-12) that are relocated<br />

from one area of the mine to another.<br />

7-2-2 Roll Crushers<br />

Roll crushers consist of two counterrotating cylinders. The gap between the cylinders is<br />

adjusted by threaded bolts. Roll crushers can use springs to hold the cylinders in place.<br />

Each cylinder is then driven by its own belt drive.<br />

Roll crushers are used for less abrasive stones than cone crushers. They are most effective<br />

on soft and friable stones, or when a close-sized product is required.<br />

7-3 GRINDING CIRCUITS<br />

COMPONENTS OF SLURRY PLANTS<br />

FIGURE 7-9 Performance curves of an impact crusher. (Courtesy of Sandvik.)<br />

The dry ore from crushers is stored in a stockpile (see Figure 1-10). The stockpile then<br />

feeds the milling circuit (Figure 7-13). It is claimed that grinding accounts for 60% of the<br />

power consumption of a mineral process plant. Elliott (1991) indicates that for a typical<br />

cop<strong>pe</strong>r or zinc concentrator, grinding consumes 12 kWh/t, crushing 2–3 kWh/t, and the<br />

rest of the plant 2–3 kWh/t. Obviously, the finer the grinding, the higher the energy consumption.<br />

There are two main forms of grinding:<br />

1. Dry grinding when the water content is 34% water by volume<br />

7.11


7.12 CHAPTER SEVEN<br />

FIGURE 7-10 Cross-sectional cut through a cone crusher.(Courtesy of Sandvik.)<br />

Between 1% and 34%, the <strong>slurry</strong> is very difficult to handle and grinding is inefficient. In<br />

some plants, an initial grinding process may be followed by some form of classification<br />

such as flotation or magnetic separation, which in turn is followed by a second grinding<br />

process. This approach tends to eliminate at an early stage a good portion of the gangue<br />

(see Chapter 1).<br />

It is not possible to achieve the particle size needed through a single grinding phase<br />

unless coarse output is required. When a coarse product is required, crushed materials are<br />

transported to a rod mill via a conveyor belt and the output is delivered from the rod mill.<br />

This is essentially an o<strong>pe</strong>n circuit.<br />

Closed circuits (Figures 7-14–7-16) may include SAG and ball mills, hydrocyclones,<br />

and centrifuges. Grinding mills are designed with different approaches to feed and discharge<br />

(Figure 7-17). The energy required to reduce the size of a particle is usually a<br />

function of its diameter raised to an exponent. Holmes (1957) indicated that this exponent


FIGURE 7-11 Gradation curves of cone crushers. (Courtesy of Sandvik.)<br />

FIGURE 7-12 Mobile cone and screen plants. (Courtesy of Sandvik.)<br />

7.13


7.14 CHAPTER SEVEN<br />

reclaim water<br />

Crushers<br />

SAG Mill<br />

SAG Mill discharge<br />

Cyclone<br />

Feed<br />

Pumps<br />

conveyor<br />

Stockpile<br />

Monorail<br />

Water<br />

Sprays<br />

cyclone overflow<br />

Ball Mill<br />

Mill Feed<br />

Conveyor<br />

coarse<br />

Water<br />

Sprays<br />

Auto<br />

Sampler<br />

reclaim<br />

water<br />

Reclaim<br />

water<br />

Mill Feed Stockpile<br />

Belt<br />

Feeders<br />

To Rougher Flotation<br />

FIGURE 7-13 Flow chart of a grinding circuit. The stockpile of ore feeds the SAG mill, and<br />

the ore is processed even further by ball mills.<br />

is not a constant but a variable. His method of iteration is fairly complex and would require<br />

a computer program.<br />

For wet grinding, which is where the <strong>slurry</strong> circuit starts, the resistance to comminution<br />

is measured by a grindability work index. It is established by test work. Bond (1952)<br />

defined the grindability work index � from the power W (in kWh <strong>pe</strong>r ton) required to reduce<br />

the feed size F (mm) to the final product size Cr (mm):<br />

–1/2 –1/2 W = 10�(Cr80 – Fe80 ) (7-1)<br />

Equation 7-1 is based on reduction of the rock size in a 2.44 m (96 in) ball mill. This<br />

equation applies in the case of wet grinding, which is often the first step in a <strong>slurry</strong> circuit.<br />

Typical examples of the grindability work index � are presented in Table 7-1.<br />

The feed, its sha<strong>pe</strong>, and mechanical pro<strong>pe</strong>rties ultimately influence the <strong>pe</strong>rformance of<br />

the grinding circuit and the degree of efficiency of ore extraction. The <strong>pe</strong>rformance of the<br />

grinding process is de<strong>pe</strong>ndent on a successful grinding o<strong>pe</strong>ration.<br />

In an autogenous mill, the feed itself is used as a grinding medium. The larger the particles,<br />

the more energy they release on impact with each other. A coarse feed (larger than


7.15<br />

Conveyor from stock pile<br />

mill feed box<br />

feed<br />

rods<br />

primary grinding mill<br />

separation of<br />

grinding medium<br />

gear<br />

mill discharge pump box<br />

cyclone overflow (fines)<br />

separation of<br />

grinding balls<br />

cyclone feed pump<br />

or mill discharge pump<br />

FIGURE 7-14 Two-stage closed circuit for grinding and classification of ore.<br />

hydrocyclone<br />

coarse cyclone underflow<br />

recirculated to ball mill<br />

ball mill<br />

feed<br />

mill feed box


FIGURE 7-15 View of a closed circuit grinding cop<strong>pe</strong>r ore. In the back of the photo is the<br />

large 12.2 m (40 ft) diameter SAG mill that receives the ore from the stockpile. In the front, the<br />

ball mill grinds the underflow from the hydrocyclone.<br />

FIGURE 7-16 View of the hydrocyclones set at a height of 30 m above the base of the SAG<br />

mill. The overflow is diverted to centrifuges to separate the gold ore from the lighter cop<strong>pe</strong>r<br />

ore. The cop<strong>pe</strong>r ore is then diverted to the ball mill (on the left-hand side of the photo) for secondary<br />

grinding.<br />

7.16


feed<br />

feed<br />

balls<br />

COMPONENTS OF SLURRY PLANTS<br />

out<br />

FIGURE 7-17 Schematic representation of different ty<strong>pe</strong>s of grinding mills.<br />

feed<br />

7.17<br />

<strong>slurry</strong> grate<br />

(a) Overflow mills (wet grinding only)<br />

(b) Diaphragm or grate mills<br />

- Used for rod mills in o<strong>pe</strong>n circuits and ball mills<br />

- Not suitable for rod mills, and mostly used for<br />

in closed circuit<br />

closed circuit<br />

Grinding with maximum s<strong>pe</strong>cific area and suitable<br />

- Used for Autogeneous and Semi-Autogeneous Grinding<br />

for very fine output<br />

for very fine output<br />

Simple and robust - Coarser output than overflow mills<br />

rods<br />

(c) <strong>pe</strong>ripheral central port discharge (d) <strong>pe</strong>ripheral discharge at the end<br />

Peripheral discharge mills are essentially reserved for rod mill grinding, wet or dry<br />

Used for coarse grind where close control of final feed size is required , either coarse or fine<br />

suitable for o<strong>pe</strong>n or closed circuits<br />

TABLE 7-1 Typical Examples of Grindability Work Indices (For Wet Grinding in a<br />

Ball Mill)<br />

Grindability<br />

Material work index Reference<br />

Barite 5 Elliott (1991)<br />

Bauxite 9 Elliott (1991)<br />

Clay 7 Elliott (1991)<br />

Coal 11 Elliott (1991)<br />

Dolomite 11 Elliott (1991)<br />

Feldspar 12 Elliott (1991)<br />

Fluorspar 9 Elliott (1991)<br />

Granite 15 Elliott (1991)<br />

Limestone 12 Elliott (1991)<br />

Magnetite 10 Elliott (1991)<br />

Quartz 13 Elliott (1991)<br />

Quartzite 10 Elliott (1991)<br />

Sandstone 7 Elliott (1991)<br />

Shale 16 Elliott (1991)<br />

Taconite 23 Elliott (1991)<br />

feed<br />

rods<br />

out<br />

feed


7.18 CHAPTER SEVEN<br />

150 mm or 6 in) is important for a fully autogenous mill. Typically, the feed has an 80%<br />

passing size of 200 mm (8 in).<br />

In a semiautogenous (SAG) mill, steel or high chrome white iron balls are added to the<br />

circuit as a grinding medium. As they rotate and are carried away by centrifugal forces,<br />

they fall by gravity and impact against the feed or crushed rocks. Due to the difference in<br />

density between the steel balls (typically 7610 kg/m3 or a s<strong>pe</strong>cific gravity of 7.61) and<br />

rocks (with a range of s<strong>pe</strong>cific gravity of 1.3 to 4.0), smaller steel balls in a SAG mill<br />

have the effect of large rocks in fully autogenous mills. The d80 of the feed, called F80 in<br />

SAG mills, is typically 110 mm (4.5 in).<br />

In a mineral process plant, the process of comminution is one of the least efficient and<br />

highest consumers of power. A number of equations are used to define the process of dry<br />

grinding. These are described by Elliott (1991).<br />

Equation 7.1 is often called Bond equation. In practice it is modified by multiplying<br />

the right hand side of the equation by so-called “inefficiency factors,” E1 to E9. Dry grinding correction factor E1. For dry grinding circuits, without the addition of<br />

water, an inefficiency factor, E1 = 1.3, is applied.<br />

Product size correction factor E2. Another efficiency factor in terms of the final<br />

product size is defined as E2. If the final product is classified at 80% of the passage diameter,<br />

then E2 = 1.2. If the final product is classified at 95%, then E2 = 1.57 (see Table 7-2).<br />

Diameter correction factor E3. For a mill with the diameter Dm (in meters), a coefficient<br />

E3 is defined as<br />

E3 = (2.44/Dm) 0.2 (7-2a)<br />

If the diameter of the mill is expressed in inches then<br />

E3 = (96/Dmus) 0.2 (7-2b)<br />

where Dmus is the diameter of the mill in inches.<br />

Oversize correction factor E4. The optimum rock size fed into a rod mill is given as<br />

Feop = 16,000 (13/�) 1/2 expressed in �m (7-3)<br />

and for a ball mill:<br />

Feop = 4000 (13/�) 1/2 expressed in �m (7-4)<br />

TABLE 7-2 Inefficiency Factor E 2 for Grinding<br />

Circuits<br />

Product size<br />

control reference % passing E 2<br />

50% 1.035<br />

60% 1.05<br />

70% 1.10<br />

80% 1.20<br />

90% 1.40<br />

92% 1.46<br />

95% 1.47<br />

98% 1.70<br />

Source: “The Science of Communition,” Brochure No.<br />

0647-05-98-N-English, Nordberg, Helsinki, Finland, 1998.


COMPONENTS OF SLURRY PLANTS<br />

If the size of the feed is larger than the optimum size Feop, (i.e., if Fe80 � Feop), then E4 =<br />

1 if Fe80 > Feop (the case of oversized feed); then<br />

Fe80 – Feopt Cr80 E4 = 1 + (� – 7)���� (7-5)<br />

Feopt*<br />

When Equation 7-5 yields a result smaller than 1.0, the result should be corrected to E4 =1.0. This equation should not be used in the case of a rod mill used to feed a ball mill, in<br />

which case, E4 = 1.0.<br />

Fineness correction factor E5. If the crushed output diameter Cr80 is less than 75 �m,<br />

then it is necessary to calculate a fineness correction factor E5, defined as<br />

Cr80 + 10.3<br />

E5 = ��<br />

(7-6)<br />

1.145Cr80<br />

Otherwise E5 = 1.<br />

Correction factor for high/low ratio of reduction rod milling E6. For a rod mill,<br />

defining the length of the mill as Lm and the diameter as Dm, a ratio Rr0 is defined as<br />

Rr0 = 8 + (5Lm/Dm) (7-7)<br />

The material reduction ratio is defined as<br />

Rr = Fe80/Cr80 (7-8)<br />

If Rr > (Rr0 ± 2), then<br />

(Rr – Rr0) E6 = 1 + � � (7-9)<br />

Otherwise a correction factor E6 = 1 is assumed.<br />

Correction factor for the low reduction ratio for ball mills. If Rr < 6, or when the<br />

ratio of the ball mill feed to the product output sizes is smaller than 6.0, a correction factor<br />

E7 is defined as<br />

2(Rr – 1.35) + 0.26<br />

E7 = ��<br />

(7-10)<br />

2(Rr – 1.35)<br />

2<br />

��<br />

150<br />

If the computation of Equation 7-10 exceeds the magnitude of 2.0, it is highly recommended<br />

to conduct lab tests and to contact the manufacturer of the mills.<br />

Correction factor for rod mills E8. The rod milling feed factor is where the material<br />

is fed into a rod mill from an o<strong>pe</strong>n circuit crusher. Elliott (1991) suggested 1.4 as the magnitude<br />

of E8. However, if the source is a closed circuit with rod milling followed by ball<br />

milling, then E8 is 1.2.<br />

Correction factor for rubber-lined mills E9. When grinding balls are smaller than<br />

80 mm or 3.25 in, rubber liners are used to line the inside walls of the mill. When grinding<br />

balls are larger than 80 mm or 3.25 in, metal liners are used.<br />

Rubber liners (Figure 7-18) are thicker than metal liners, use more space, and absorb<br />

more impact energy than their metal counterparts. It is customary to apply a correction<br />

factor E9 = 1.07 for rubber liners.<br />

The final power required to mill the feed is then obtained after multiplying all the correction<br />

factors by Bond’s equation (7-1).<br />

Iteration to consumed energy:<br />

Wf = W(E1 × E2 × E3 × E4 × E5 × E6 × E7 × E8 × E9) (7-11)<br />

� Fe80<br />

7.19


7.20 CHAPTER SEVEN<br />

FIGURE 7-18 Rubber lining of SAG mills supplied to the Murin–Murin project in Australia<br />

to treat nickel-rich laterites. [Courtesy of Metso Minerals (formerly known as the companies<br />

Nordberg and Svedala).]<br />

Equation 7-11 is useful to determine the power to grind down rocks. It must be corrected<br />

for worn-out liners, ball charges, and <strong>slurry</strong> density. It is therefore recommended that in<br />

the initial phase of the design of a mineral process plant, lab tests be conducted.<br />

Some of the empirical coefficients and equations for E 1 to E 9 were develo<strong>pe</strong>d assuming<br />

a recirculation load of 250%. This means that the charge load of coarse material that<br />

is returned to the mill is about 250% of the fresh feed in a closed circuit. This is not always<br />

the case. The author was once involved in the design of a cop<strong>pe</strong>r concentrate plant<br />

for a Peruvian mine in which the presence of soft high clay in the ore increased viscosity<br />

tremendously at a weight concentration of 50% to 60%. It became necessary to add water,<br />

dilute the <strong>slurry</strong>, and cut down the recirculation load.<br />

When the rocks in the feed are large, and milling is dominated by impact loads, Equation<br />

7.1 should not be used to compute the work index load.<br />

Some of the empirical coefficients and equations for E 1 to E 9 were develo<strong>pe</strong>d for a final<br />

output size with 80% passing 100 �m. (mesh 140). When C r80 < 100 �m, Equation<br />

7.11 does not give correct results.<br />

Example 7-1<br />

An ore with a grindability index � = 13 is to be ground in a rod mill with feed from a<br />

closed-circuit crusher. The feed has a diameter F e80 of 26 mm (1 in). The final product is<br />

required at 80% to be C r80 of 10 mm (0.4 in) at a mass throughput of 350 tons/hour<br />

(770,000 lbs/hour). Estimate the power consumed by the rod mill.


Solution<br />

Using Equation 7-1, the work input to the rod mill is<br />

W = 10 × 13(10 –1/2 – 26 –1/2 ) = 130(0.3162 – 0.1961) = 15.61 kWh/ton<br />

For wet grinding, E1 = 1. For closed-circuit grinding E2 = 1; E3 will be calculated after<br />

other factors.<br />

The oversize feed factor E4 is obtained from Equation 7.3.<br />

Feop = 16,000(13/13) 1/2 = 16,000 �m or 16 mm<br />

Since Feop < Fe80, then<br />

E4 = {[(26/10) + (13 – 7)(26 – 16)]/16}/(26/10) = 0.3846(2.6 + 3.75) = 2.442<br />

Since Cr80 > 75 �m, then E5 = 1.<br />

From Equation 7-8, the reduction ratio of the material Rr = 26/10 = 2.6. Rr0 will be calculated<br />

after selecting the rod mill. Since Rr < 6 then<br />

E7 = [2(2.6 – 1.35) + 0.26]/2(2.6 – 1.35) = 1.104<br />

E8 = 1.2 since it is a closed circuit crusher.<br />

Iteration to consumed energy<br />

Wf = W(E1 × E2 × E3 × E4 × E5 × E6 × E7 × E8) Wf = 15.61 × 1 × 1 × E3 × 2.442 × 1 × E6 × 1.104 × 1.2 = 50.5 × E3 × E6 kWh/ton<br />

Since the feed is 350 tons <strong>pe</strong>r hour, the total energy consumption would be<br />

350 ton/h × 50.5 kWh/ton E3 × E6 = 17,675 kW × E3 × E6 This would require a number of mills in parallel. From Equation 7-2, if the mill diameter<br />

of 6 m (19.7 ft) is selected, then<br />

E3 = (2.44/6) 0.2 = 0.833<br />

Rod mills with a length to diameter ratio of 2 are selected:<br />

Rr0 = 18<br />

and since Rr < (Rr0 ± 2),<br />

E6 = 1<br />

Final power consumption is 42.067 kWh/ton or total of 14,723 kW (19,736 hp).<br />

With modern technology, a SAG mill should be considered as an alternative to the rod<br />

mill (see Tables 7-3 and 7-4).<br />

7-3-1 Single-Stage Circuits<br />

COMPONENTS OF SLURRY PLANTS<br />

7.21<br />

When finer material is required, a ball mill is used in a closed circuit. The feed is ground<br />

and then classified to separate coarse from fine solids. The coarse solids, also called oversized<br />

particles, are returned back to the mill for further grinding. This is called the “recirculation<br />

load” and the circuit is considered a closed circuit. In a dry circuit, the classifier<br />

may be a set of vibrating screens.<br />

In a typical cop<strong>pe</strong>r or zinc circuit, the recirculation load can be as high as 250–350%<br />

of the new feed. The mill and mill discharge pumps must then be sized for the combination<br />

of recirculation load and new feed.


7.22 CHAPTER SEVEN<br />

TABLE 7-3 Estimates of Bond Energy Consumption <strong>pe</strong>r Mass for Grinding<br />

Rocks (W i)<br />

Mineral S<strong>pe</strong>cific gravity (kWh/sh.ton) (kWh/tonne)<br />

Andesite 2.84 18.25 20.08<br />

Barite 4.50 4.73 5.20<br />

Basalt 2.91 17.10 18.81<br />

Bauxite 2.20 8.78 9.66<br />

Cement clinker 3.15 13.45 14.80<br />

Clay 2.51 6.30 6.93<br />

Coal 1.4 13 14.3<br />

Coke 1.31 15.13 16.84<br />

Cop<strong>pe</strong>r ore 3.02 12.72 13.99<br />

Diorite 2.82 20.90 22.99<br />

Dolomite 2.74 11.27 12.40<br />

Emery 3.48 56.70 62.45<br />

Feldspar 2.59 10.80 11.88<br />

Ferro-chrome 6.66 7.64 8.40<br />

Ferro-manganese 6.32 8.30 9.13<br />

Ferro-silicon 4.41 10.01 11<br />

Flint 2.65 26.16 28.78<br />

Fluospar 3.01 8.91 9.8<br />

Gabbro 2.83 18.45 20.3<br />

Glass 2.58 12.31 13.54<br />

Gneiss 2.71 20.13 22.14<br />

Gold ore 2.81 14.93 16.42<br />

Granite 2.66 15.13 16.64<br />

Graphite 1.75 43.56 47.92<br />

Gravel 2.66 16.06 17.67<br />

Gypsum rock 2.69 6.73 7.40<br />

Iron ore, hematite 3.53 12.84 14.12<br />

Iron ore, hematite—s<strong>pe</strong>cular 3.28 13.84 15.22<br />

Iron ore, magnetite 3.88 9.97 10.97<br />

Iron ore, oolitic 3.52 11.33 12.46<br />

Iron ore, taconite 3.54 14.61 16.07<br />

Lead ore 3.35 11.90 13.09<br />

Lead–zinc ore 3.36 10.93 12.02<br />

Limestone 2.66 12.74 14<br />

Manganese ore 3.53 12.20 13.42<br />

Magnesite 3.06 11.13 12.24<br />

Molybdenum 2.70 12.80 14.08<br />

Nickel ore 3.28 13.65 15.02<br />

Oil shale 1.84 15.84 17.43<br />

Phosphate rock 2.74 9.92 10.91<br />

Potash ore 2.40 8.05 8.86<br />

Pyrite ore 4.06 8.93 9.83<br />

Pyrhotite ore 4.04 9.57 10.53<br />

Quartzite 2.68 9.58 10.54<br />

Quartz 2.65 13.57 14.93<br />

Rutile ore 2.80 12.68 13.95<br />

W i<br />

W i


TABLE 7-3 Continued<br />

7-3-2 Double-Stage Circuits<br />

A rod mill in an o<strong>pe</strong>n circuit may be followed by a ball mill in a closed circuit. This is<br />

called a double-stage circuit and is often a wet process. The output from the rod mills is a<br />

<strong>slurry</strong> that contains a high proportion of coarse stones. The <strong>slurry</strong> is pum<strong>pe</strong>d via “mill discharge<br />

pumps” to a hydrocyclone. The underflow from the cyclone is then fed to a ball<br />

mill. From there, the output from the ball mill is fed once again to the hydrocyclone via<br />

the pump.<br />

In some circuits, the rod mill discharge is fed first to the ball mill before reaching the<br />

hydrocyclone. The hydrocyclones then feed the ball mills by gravity. A set of ball mill<br />

discharge pumps may then pump the output to a second classification circuit. The ball<br />

mill discharge has its own sets of <strong>slurry</strong> pumps.<br />

7-4 HORIZONTAL TUMBLING MILLS<br />

In a horizontal tumbling mill, the actual body of the mill rotates and imparts energy to the<br />

grinding medium (balls or rods) and to the <strong>slurry</strong>. The combination of centrifugal forces<br />

and gravity forces from falling media act to create energy transmission by impact against<br />

the mineral. There are three categories of horizontal tumbling mills:<br />

1. rod mills<br />

2. ball mills<br />

3. autogenous and semi-autogenous mills<br />

COMPONENTS OF SLURRY PLANTS<br />

Mineral S<strong>pe</strong>cific gravity (kWh/sh.ton) (kWh/tonne)<br />

Shale 2.63 15.87 17.46<br />

Silica sand 2.67 14.10 15.51<br />

Silicon carbide 2.75 25.87 28.46<br />

Slag 2.74 10.24 11.26<br />

Slate 2.57 14.30 15.73<br />

Sodium silicate 2.10 13.40 14.74<br />

Spodumene ore 2.79 10.37 11.41<br />

Syenite 2.73 13.13 14.44<br />

Tin ore 3.95 10.90 11.99<br />

Titanium ore 4.01 12.33 13.56<br />

Trap rock 2.87 19.32 21.25<br />

Zinc ore 3.64 11.56 12.72<br />

From Denver Sala Basic. Selection Guide for Process Equipment. Reproduced by <strong>pe</strong>rmission of<br />

Metso Minerals (formerly known as the companies Nordberg and Svedala).<br />

7.23<br />

Basically a horizontal tumbling mill is a cylinder lined on the inside with wear-resistant<br />

alloy liners. The liners are fixed to the shell by T-bolts and nuts on the outside. The<br />

cylinder is carried by hollow trunnions running side bearings at each end.<br />

W i<br />

W i


7.24<br />

TABLE 7-4 Selection Guide for Grinding Mills<br />

Rod<br />

__________________<br />

Autogenous<br />

Ball Vertical<br />

_________________ Peripheral ______________________ _____________<br />

Mineral Primary Secondary Overflow Discharge Overflow Grate Pebble Spindle Tower Vibrating Hammer<br />

Ores (ferrous and nonferrous) � � � � � � � � � �<br />

Preponderance of fine aggregates � � � �<br />

Talc and ceramic materials �<br />

Cement raw materials � � �<br />

Cement clinker �<br />

Coal and <strong>pe</strong>trol, coke � � � �<br />

Silica ceramics, etc. (must be free of iron) � �<br />

Production to a s<strong>pe</strong>cific particle diameter or mesh � � � � � � � � � � �<br />

Production to a s<strong>pe</strong>cific surface area � �<br />

Wet grinding � � � � � � � � � �<br />

Dry grinding � � � � � � � �<br />

Damp feed (1%–15% moisture) �<br />

Large feed (


COMPONENTS OF SLURRY PLANTS<br />

7.25<br />

A s<strong>pe</strong>cial chamber on the tip is installed to feed the material and the grinding medium.<br />

The liners on the inside are designed to be ribbed either along the length of the cylinder or<br />

in spiral sha<strong>pe</strong>d ribs. Ni-hard is the most commonly used alloy for liners (Figure 7-19),<br />

but manganese steel and chrome steel are also used. In some designs, rubber is used as the<br />

liner material (Figure 7-18)<br />

It is important to cut down maintenance costs and <strong>pe</strong>riods of outage to replace<br />

liners. Many mines o<strong>pe</strong>n their mills once a month for maintenance and to replace the<br />

worn liners. The grinding medium is steel rods in the case of rod mills and steel balls<br />

in the case of ball mills. As the cylinder or tumbler rotates, the heavy rods and balls<br />

are lifted by the ribs of the liners. The rods and balls fall by gravity after a certain angle<br />

of rotation is reached. The impact in turn fractures and grinds the rocks into smaller<br />

stones.<br />

The spacing between the ribs of the liner is critical. Too narrowly spaced ribs may<br />

jam the coarser rocks and delay their fracture. S<strong>pe</strong>ed of rotation is extremely important.<br />

At a certain s<strong>pe</strong>ed, the material, which is lifted by friction against the liner, starts to fall<br />

down. The cascading effects of stone against stone causes grinding by attrition. The material<br />

output is fine but wear is high. As the s<strong>pe</strong>ed of the mill is increased, grinding<br />

takes place by impact of the rods or balls against the rocks. As the s<strong>pe</strong>ed increases even<br />

further, centrifugal forces become sufficient for the material to centrifuge. This s<strong>pe</strong>ed is<br />

called “the critical s<strong>pe</strong>ed of the mill.” Mills are designed to o<strong>pe</strong>rate at 75% of their critical<br />

s<strong>pe</strong>ed.<br />

The diameter of the rods is often 50 mm (2 in) but can be set by the designer of the<br />

mill. It is, however, important to separate the rods or balls from the <strong>slurry</strong> at the discharge<br />

of the mill before they enter the <strong>slurry</strong> pump. The successful separation of steel balls from<br />

the <strong>slurry</strong> involves pro<strong>pe</strong>r design of trommels, a mechanism to catch the balls, and<br />

screens on top of the pump box. Ideally, the balls should be recycled back to the feed of<br />

the milling unit.<br />

FIGURE 7-19 Worn-out metal liners removed during monthly maintenance of a SAG mill.


feed<br />

7.26 CHAPTER SEVEN<br />

7-4-1 Rod Mills<br />

Rod mills are a ty<strong>pe</strong> of fine crusher and can reduce the size of rocks down to 1 mm (0.04<br />

in). They <strong>pe</strong>rform better than a fine crusher in less than optimum conditions when the<br />

feed is damp or contains clay.<br />

Typically, the length to diameter ratio of the rod mills is 1.5 to 2.5. Milling occurs by<br />

impact of rod against rod. The stones are trap<strong>pe</strong>d between the rods and disintegrate. The<br />

coarser stones are the first to break. The finer esca<strong>pe</strong> milling. Rod mills are not used on<br />

closed circuits.<br />

In the last 20 years, the mining industry has tended to replace rod mills with large autogenous<br />

and semiautogenous mills<br />

7-4-2 Ball Mills<br />

In ball mills, metal balls are used as the grinding media. The balls are made of a variety of<br />

materials. Steel balls are forged. High chrome balls are cast with 28% chrome and are<br />

available from s<strong>pe</strong>cial foundries. About 1 kg of balls is used <strong>pe</strong>r ton of stone. Small balls<br />

with a diameter of about 25 mm (1 in) are preferred to larger ones in order to maximize<br />

the area of contact between balls and stones.<br />

The <strong>slurry</strong> weight concentration in a ball mill is 65–80%. Excessive concentration will<br />

cause the particles to stick to the balls and will decrease the effectiveness of grinding. The<br />

ball mill may then “freeze” and spill out its contents, causing costly downtime to empty<br />

the mill. For this reason, the weight concentration should not be allowed to exceed 80%.<br />

A trunnion at the discharge of the ball mill separates balls from <strong>slurry</strong>. The balls are<br />

then conveyed back to the feed. Balls gradually wear out through re<strong>pe</strong>ated feeding to the<br />

mill and must be replaced.<br />

Ball mills are built in different diameters up to a maximum of 6.5 m (21 ft), and in<br />

power drives up to 9650 kW (13,000 hp). Their sha<strong>pe</strong> is determined by the ty<strong>pe</strong> of output<br />

(Figure 7-20).<br />

7-4-3 Autogenous and Semiautogenous Mills<br />

Autogenous and semiautogenous (AG and SAG) mills are extremely large mills with a<br />

maximum diameter of 12.2 m (40 ft). In the last few years, Siemens and ABB have devel-<br />

balls<br />

<strong>slurry</strong><br />

discharge<br />

Cascade mills (wet and dry grinding)<br />

- Used for autogneous and semiautogeneous milling<br />

in closed circuit<br />

- Primary Grinding with minimum retention time<br />

for very fine output<br />

- Diameter to length ratio 2:1<br />

balls<br />

feed<br />

<strong>slurry</strong><br />

(b) Conical sha<strong>pe</strong> mill<br />

- Suitable for fine discharge<br />

discharge<br />

FIGURE 7-20 The sha<strong>pe</strong> of ball grinding mills is determined by the ty<strong>pe</strong> of discharge and ore.


o<strong>pe</strong>d and installed “wraparound” motors. In this design, the outside diameter of the tumble<br />

on one side is part of the rotor of the motor (Figure 7-21). These motors are manufactured<br />

up to a power size of 7000 kW (9400 hp).<br />

The large diameter of these mills maximizes the impact forces. Although the feed is<br />

typically 150–180 mm (6–7 in) in diameter, the output can be as fine as 0.3 mm (0.012<br />

in). Particles tend to cleave along their natural grain boundaries.<br />

Six to ten <strong>pe</strong>rcent of steel balls are added on a continuous basis to the feed to assist<br />

grinding through a separate entry. Wet milling and grinding is less dusty and less noisy<br />

than dry grinding. The feed and output trunnions are on opposite sides. The trommel on<br />

one side catches the steel or high chrome balls to prevent them from falling into the pump<br />

box.<br />

7-5 AGITATED GRINDING<br />

COMPONENTS OF SLURRY PLANTS<br />

7.27<br />

FIGURE 7-21 SAG mill with wrap-around or ring motor. [Courtesy of Metso Minerals (formerly<br />

known as the companies Nordberg and Svedala).]<br />

Agitation is another method of grinding. The whole body of the mill may sit on springs<br />

and be agitated by crankshafts or an eccentric mechanism driven by a motor. Another ap-


7.28 CHAPTER SEVEN<br />

proach is to have a rotor, an agitator, and a rotating hammer inside the mill to impart energy.<br />

7-5-1 Vertical Tower Mills<br />

A vertical mill was develo<strong>pe</strong>d in Japan for grinding fine minerals. The tower mill is a<br />

combination of vertical tanks, a screw mixer, a mill, and a classifier. From a chute on the<br />

side of the mill, the rocks, steel balls, and liquid are introduced. The vertical screw rotates<br />

around a vertical shaft and creates an upward vertical counterflow. The finer materials<br />

float to the top and are led to a side chamber for classification, while the heavier and<br />

coarser solids sink with the steel balls (Figure 7-22).<br />

The diameter of the balls is 6–32 mm (0.25–1.25 in). Size reduction of the ore is limited<br />

to 5 mm (0.197 in) due to limited grinding. Vertical tower mills are manufactured for a<br />

maximum output of 100 tph. They require limited floor space and have a low consumption<br />

of power.<br />

7-5-2 Vertical Spindle Mills<br />

The vertical spindle mill (also known as SAM) uses a central vertical multistage mixer<br />

(Figure 7-23). Each stage consists of a number of wolfram carbide pins fixed on a hollow<br />

shaft. They provide horizontal stirring. This machine o<strong>pe</strong>rates with feed smaller than 1<br />

mm (16 mesh) for fine and ultrafine wet or dry grinding. The units are small and compact<br />

and can be relocated within the plant. Maximum power is 75 kW <strong>pe</strong>r unit.<br />

7-5-3 Roller Mills<br />

Roller mills are used for soft grinding of industrial minerals in a dry state. The mill consists<br />

of a rotating table on a vertical axis. Two rollers rotate around their own shafts at an<br />

angle with res<strong>pe</strong>ct to each other. The rollers are spring loaded. The output is diverted to<br />

dry cyclones and the oversized material is fed back to the roller mill.<br />

A new generation of high-pressure roller mills has ap<strong>pe</strong>ared on the market since the<br />

1980s. A very high level of torque is transmitted to the rollers to maximizing the crushing<br />

loads. High-pressure rollers are mainly used in cement plants, diamond processing (when<br />

the extraction is from rocks, as it is in Canada), and to a certain extent in the field of metalliferrous<br />

minerals.<br />

7-5-4 Vibrating Ball Mills<br />

The body of the mill consists of a central feed chamber and two side chambers. The feed is<br />

from the top and the discharge from the central chamber is at the opposite end. The whole<br />

body of the mill sits on four strong springs. Two electric motors synchronized by V-belts<br />

rotate an eccentric mechanism linked to each of the side chambers (Figure 7-24). This machines<br />

uses fine feed smaller than 5 mm (mesh 4) and is particularly suited for difficult material<br />

with an energy index W i > 30 kWh/sh.ton. These are essentially small machines with<br />

maximum motor size rated at 55 kW or 75 hp. However, they are often chosen over tumbling<br />

mills for lower installation cost, lower o<strong>pe</strong>rating cost, less floor space, increased<br />

grinding flexibility, and improved product control within the limitation of their size. Rods<br />

or balls may be used as grinding media within o<strong>pe</strong>n or closed circuits with these machines.


7.29<br />

�������� ���� ��� ����<br />

���������� ��� ���� �� ��������<br />

����� ��� ������� �������<br />

����� ������ ��� ��������<br />

������ ����<br />

����� ����<br />

���������� ���<br />

������� ��� ����� ��������<br />

������������� �� ������ ��������<br />

������ ����<br />

FIGURE 7-22 Slurry circuit of vertical grinding tower mill for solids with a maximum diameter of 6.4 mm (1/4�).


7.30<br />

vertical bars for wall<br />

protection and help to grinding<br />

helix for upwards pumping<br />

while mixing and grinding<br />

solids feed classifier box<br />

FIGURE 7-23 Vertical spindle mill <strong>slurry</strong> circuit.<br />

launder for fines (output)<br />

recirculation of coarse material<br />

M<br />

A<br />

A<br />

Z<br />

D<br />

P U -<br />

A<br />

K<br />

2 5<br />

<strong>slurry</strong> pump


two motors in parallel at<br />

each end<br />

7-5-5 Hammer Mills<br />

In the hammer mill, a central rotor arm is fitted with rings of arms that crush and mill the<br />

feed against the wall of the mill. It is essentially used for dry milling of low-abrasive and<br />

friable minerals such as cement, coal, gypsum, and limestone. It is considered by some<br />

engineers a crusher rather than a mill.<br />

7-6 SCREENING DEVICES<br />

COMPONENTS OF SLURRY PLANTS<br />

feed<br />

one side chamber at each end<br />

discharge<br />

FIGURE 7-24 Vibrating mill.<br />

7.31<br />

In Chapters 1 and 3, the concept of d 50 was introduced as the particle size diameter below<br />

and at which 50% of the particles can pass through the o<strong>pe</strong>ning of a sieve. The<br />

same concept applies to the definition of screen size. The screen size a<strong>pe</strong>rture is equal<br />

to d 50.<br />

An ideal screen would let all particles equal to or smaller than d 50 pass through. This is<br />

not always the case, as the <strong>pe</strong>rformance of the screen de<strong>pe</strong>nds on a variety of factors:<br />

� Screen deck size. In order for all particles to use the screen effectively, the layer of<br />

solids above the screen needs to be very thin. This means a large deck size for a given<br />

mass of solids. For economical reasons, this is not possible and a thick layer of solids<br />

forms on the smaller screens.<br />

� Vibration. To move away the coarse particles that block the passage of the finer ones,<br />

it is essential to oscillate the screen. The amplitude of the oscillation must match the<br />

s<strong>pe</strong>cifics of the solids. Too much vibration could cause the solids to float as a cloud<br />

without passing through the screens.<br />

� Presentation angle. Ideally, the solids should be fed as normal as possible to the screen.<br />

This means that the solids should come in at a 90° angle. Unfortunately, this is not always<br />

possible.<br />

� Screen material. Screens are manufactured of metal, rubber, and even fiberglass. Metal<br />

screens have a wider a<strong>pe</strong>rture than rubber, which is more flexible and less prone to particle<br />

binding.<br />

� Moisture content. Sprays are sometimes added to screens to improve their efficiency<br />

and flush the solids. Sprays suppress clouds of fine particles.


7.32 CHAPTER SEVEN<br />

7-6-1 Trommel Screens<br />

Trommel screens are essentially rotating cylindrical mesh. These trommels rotate at a<br />

slight angle of inclination to facilitate the removal of material. Trommel screens can be<br />

supplied as a set of concentric screens of different a<strong>pe</strong>rture. The finest screens are the<br />

quickest to show wear.<br />

7-6-2 Shaking Screens<br />

Shaking screens move in a horizontal reciprocal motion along the length of the screen.<br />

Solids are fed in a horizontal circular movement. The discharge moves in the direction of<br />

the horizontal movement of the screen. The motion of the solids changes from circular at<br />

the feed to eccentric, and finally to horizontal shaking.<br />

7-6-3 Vibrating Screens<br />

These screens are set at an angle with res<strong>pe</strong>ct to the horizontal. The vibration occurs at a<br />

right angle to the screen by the rotation of unbalanced counterweights on a shaft above<br />

the screen. Vibration can also be induced by electromagnets and oscillating currents. Vibration<br />

levels are high and the screens must be mounted on vibration isolation rubber<br />

pads. These screens are extremely noisy and exceed 100 dBA levels of noise, nevertheless,<br />

they are the most widely used.<br />

7-6-4 Banana Screens<br />

Banana screens are essentially stationary screens. The sieve is bent around a curved<br />

screen. The top of the screen is vertical and solids are fed from the top. Particles pass successive<br />

wedge bars and solids are removed between them based on the trigonometric<br />

o<strong>pe</strong>ning normal to the fall. To avoid clogging, the bars are pneumatically tilted at regular<br />

intervals. These screens can be designed to sieve particles as fine as 50 �m.<br />

7-7 SLURRY CLASSIFIERS<br />

Classification is the process that separates coarse from fine. Various methods use the effect<br />

of size, density, and magnetic and electrostatic pro<strong>pe</strong>rties of the solids. When the<br />

weight concentration is smaller than 15%, particles settle in a “free settling” mode. When<br />

the weight concentration increases, turbulence promotes settling of the heavier particles<br />

faster than the lighter particles. Two families of density classifiers are available:<br />

1. Classifiers that use the principle of free settling to achieve size separation<br />

2. Classifiers that use the particles’ hindered settling s<strong>pe</strong>ed for density separation and for<br />

concentration of a particular mineral<br />

7-7-1 Hydraulic Classifiers<br />

In a hydraulic classifier, solids are fed at the top through a chamber that leads into a column.<br />

Water is pum<strong>pe</strong>d from the bottom of the column. The counterflow moves into a


number of successive columns or stages. Products settle in accordance with the principles<br />

described in Chapter 3. Solids are removed at the bottom through a restriction such as a<br />

spigot. The spigot valve o<strong>pe</strong>ns and closes in accordance with the applied pressure from<br />

the accumulation of solids. This is the principle of o<strong>pe</strong>ration of the hydrosizer, which is<br />

often connected to other gravity and magnetic separators.<br />

7-7-2 Mechanical Classifiers<br />

A mechanical classifier is a combination of an o<strong>pe</strong>n channel flow, a weir, and a mechanical<br />

device to remove solids. The trough to which the <strong>slurry</strong> is directed is inclined<br />

with res<strong>pe</strong>ct to the horizontal. On one side, a weir is constructed opposite to the direction<br />

of the flow. The heavier solids deposit upstream of the weir, whereas the finer<br />

solids pass over it. This system o<strong>pe</strong>rates on the principle of the sliding bed described in<br />

Chapters 4 and 6.<br />

The weir is designed to minimize turbulence. A mechanical rake or rotating<br />

Archimedean screw removes the coarse solids. Water drains away as the solids are removed.<br />

The s<strong>pe</strong>ed of the rake or rotating shaft is critical to the efficiency of separation.<br />

The height of the weir must be adjustable to change the depth of the pool, the rising velocity,<br />

and the cut point between coarse and fine.<br />

The addition of water controls the density of the <strong>slurry</strong>. Dilution may be required for<br />

effective separation, but excessive water dilution may have to be followed by thickening<br />

after classification.<br />

Mechanical classifiers are ex<strong>pe</strong>nsive to install but in some applications they are selected<br />

for high-density valuable minerals as they assist in their immediate recovery without<br />

requiring further complicated flotation circuits. In some res<strong>pe</strong>cts, flotation circuits use<br />

some of the principles of mechanical classifiers by using a circular internal weir, a mixer,<br />

an underflow pump, and a separate froth pump.<br />

7-7-3 Hydrocyclones<br />

COMPONENTS OF SLURRY PLANTS<br />

Hydrocyclones (Figure 7-25) are the most common classifiers in the mining industry.<br />

They require little space and o<strong>pe</strong>rate on the pressure from the mill discharge pumps (typically<br />

104–152 kPa, 15–22 psi). They are typically used to classify solids from a size of 40<br />

to 400 �m (mesh 325 to 35).<br />

The principle of the cyclone relies on creation of a vortex; sometimes primary and secondary<br />

vortexes are created by feeding the material tangentially. In a vortex, a certain<br />

pressure field is created to counterbalance centrifugal forces. In the cyclone, the applied<br />

pressure is converted into a swirling motion. The intensity of the swirl is measured as the<br />

swirl number:<br />

S =<br />

angular momentum<br />

���<br />

axial momentum<br />

7.33<br />

If the swirl flow number exceeds 0.5, the swirl is classified as a strong swirl. Strong swirl<br />

is associated with a low-pressure zone at the core. A strong swirl is associated with an important<br />

pressure drop. The cyclone feed gauge pressure at the inlet flange is often of the<br />

order of 70–100 kPa (10 to 15 psi).<br />

The swirling chamber of the cyclone is where the separation starts. The inlet area to<br />

the cyclone is often of the order of 5–7% of the chamber area, or the inlet diameter is be


7.34 CHAPTER SEVEN<br />

FIGURE 7-25 A number of hydrocyclones may be used in parallel to classify <strong>slurry</strong> flow in<br />

a grinding circuit.<br />

tween 20% and 25% of the swirling chamber diameter. The inlet nozzle is rectangular in<br />

sha<strong>pe</strong> in some sheet metal fabricated cyclones, or circular in fiberglass cyclones.<br />

Inside the swirling chamber, a pi<strong>pe</strong> section protrudes from the top of the cyclone. It is<br />

called the vortex finder. It must extend below the feed entrance to avoid shortcuts of unclassified<br />

<strong>slurry</strong> to the top discharge or overflow. The diameter of the vortex finder is typically<br />

32% to 36% of the swirling chamber diameter. The finer and lighter particles flow<br />

out of the hydrocyclone through the vortex finder (Figure 7-26).<br />

In some of the earlier metal fabricated designs, the swirling chamber consisted of a<br />

single cylinder. In fiberglass designs, it is split into two halves, which are individually<br />

lined with removable rubber liners. The cyclone chamber is followed by a cylindrical<br />

chamber with a depth approximately equal to its diameter. This chamber provides some<br />

retention time.<br />

The cylindrical transition chamber is followed by a conical chamber, often designed<br />

with an included angle between 10 and 20 degrees. It provides further retention time.<br />

At the bottom of the cyclone, the a<strong>pe</strong>x is installed. It acts as a sort of nozzle or orifice.<br />

For different applications, different orifice diameters may be used, and for different a<strong>pe</strong>x<br />

diameters, different pressures are required. The a<strong>pe</strong>x is therefore a sort of controlling element<br />

to the cyclone. The minimum a<strong>pe</strong>x orifice diameter is on the order of 10% of the<br />

swirling chamber diameter, and the largest orifice diameter is on the order of 35%. In either<br />

case, the a<strong>pe</strong>x must allow the flow of the coarse materials. At the bottom of the a<strong>pe</strong>x,<br />

the discharge is called the cyclone underflow. At the top of the vortex finder, the discharge,<br />

which consists of fines, is called the cyclone overflow.<br />

For primary grinding circuits, the underflow typically contains 50 to 53% by volume


Involute design<br />

inlet pi<strong>pe</strong><br />

COMPONENTS OF SLURRY PLANTS<br />

vortex finder<br />

fiberglass body<br />

discharge pi<strong>pe</strong><br />

for coarse solids<br />

(cyclone underflow)<br />

discharge of finer<br />

particles (cyclone overflow)<br />

cylindrical feed<br />

chamber (top half)<br />

cylindrical chamber<br />

(bottom half)<br />

7.35<br />

Conical section<br />

angle from 10 to 20 degrees<br />

Rubber liner<br />

A<strong>pe</strong>x<br />

FIGURE 7-26 A cross-sectional representation of a rubber-lined cyclone. (Courtesy of Mazdak<br />

International Inc.)<br />

of solids, whereas for regrinding circuit, the underflow typically delivers 40 to 45% solids<br />

by volume.<br />

Hydrocyclones can be manufactured from dough-molded compound fiberglass, cast<br />

iron, or sheet metal lined with polyurethane. Metal and fiberglass cyclones are lined with<br />

rubber or with hard metal (Ni-hard or 28% chrome white iron). Burgess and Abulnaga<br />

(1991) presented a finite element analysis of fiberglass cyclones.<br />

The <strong>pe</strong>rformance of the hydrocyclone is calculated by using a partition curve similar<br />

to a screen curve. This gives the d 50 size, or 50% probability at the cut point. This cut<br />

point is defined as the condition for which 50% of the feed will be discharged as coarse<br />

particles in the cyclone underflow and 50% as fines or cyclone overflow. For every cyclone<br />

design, there is a base d 50C or cut-off for the recovery (Figure 7-27).<br />

Cyclones are usually o<strong>pe</strong>rated in a steady mode with constant pressure. Surges can<br />

lead to unfavorable air entrainment. To maintain constant pressure from the pumps, the<br />

pump box must have a constant level of <strong>slurry</strong>. To adjust the <strong>slurry</strong> level, the sump must<br />

be provided with a water addition mechanism.<br />

Normal feed to cyclones consists of a <strong>slurry</strong> at 30% solids concentration by weight.<br />

Some mines o<strong>pe</strong>rate with <strong>slurry</strong> weight concentrations as high as 35%. Higher concentration<br />

by weight imposes higher pressures of o<strong>pe</strong>ration, which can cause a reduction in<br />

efficiency of o<strong>pe</strong>ration of the hydrocyclone while coarsening the cut point. Vortex finders<br />

are changed in accordance with the required cut. A larger diameter vortex finder<br />

tends to coarsen the overflow while increasing its discharge flow rate at a constant pressure.


7.36 CHAPTER SEVEN<br />

100<br />

50<br />

Recovery to underflow, %<br />

0<br />

ACTUAL RECOVERY<br />

d<br />

Diameter of particles in micrometers<br />

The size of the discharge spigot is selected in order to maximize the flow of solids<br />

at high density and to reduce the flow of water in the underflow. Too small a spigot<br />

tends to dewater the underflow and to break the air core while reducing the overall efficiency.<br />

Because of the wide range of slurries with different particle sizes, the cutoff d50C is<br />

used to normalize the particle size (Arterburn, 1982). The actual particle diameter from a<br />

recovery is divided by the d50C size and a parameter X is defined as:<br />

X = particle diameter/d50C particle diameter<br />

The recovery to the underflow (Arterburn, 19xx) on a corrected basis is defined as:<br />

Rr = (7-15)<br />

Using the base d50C, Arterburn (1982) proposed to use three correction factors for an application,<br />

C1, C2, C3, or<br />

d50C(application) = d50C(base) × C1 × C2 × C3 (7-16)<br />

The base d50C is defined as a polynomial function of the cyclone swirling chamber diameter.<br />

Arterburn proposed<br />

d50C = 2.84(D/100) 0.66 e<br />

(7-17)<br />

where D is the cyclone chamber diameter in meters.<br />

The correction factor C1 is based on the volumetric concentration of solids fed to the<br />

cyclone:<br />

4X – 1<br />

��<br />

4X 4 e + e – 2<br />

50c<br />

CORRECTED RECOVERY<br />

FIGURE 7-27 Typical particle recovery curve for the underflow from a hydrocyclone.


COMPONENTS OF SLURRY PLANTS<br />

7.37<br />

C1 = � � –1.43<br />

(7-18)<br />

Equation 7-18 applies in a range CV < 0.4.<br />

The pressure drop between the feed nozzle and the cyclone overflow �P is used to<br />

compute the second correction factor C2: C2 = 3.27(�P) –0.28 53 – 100CV ��<br />

53<br />

(7-19)<br />

where �P is expressed in kPa. To minimize energy losses, �P should be in the range of<br />

40 to 70 kPa, Arterburn (1982), particularly for coarse separation in grinding circuits.<br />

The final correction factor C3 is based on the density of the solids with res<strong>pe</strong>ct to the<br />

liquid density:<br />

1650<br />

C3 = � (7-20)<br />

�� �S – �L Tables 7-5 and 7-6 show the typical size ranges for cyclones.<br />

Example 7-2<br />

A hydrocyclone with a diameter of 250 mm is selected for a flow rate of 15 L/s at a pressure<br />

of 140 kPa. The s<strong>pe</strong>cific gravity of the solids is 4.8 and the volumetric concentration<br />

is 0.3. If the pressure drop between the feed and the overflow is maintained at 50 kPa, determine<br />

the corrected d50C. Assuming a discharge coefficient of 0.5 and a remaining pressure<br />

of 20 kPa at the a<strong>pe</strong>x, determine the underflow capacity for an a<strong>pe</strong>x diameter of 80<br />

mm if the underflow density is 2000 kg/m3 .<br />

From Equation 7-17 the base d50C = 2.84 (D/100) 0.66 = 23.77 �m. From Equation 7-18<br />

the correction factor C1 is<br />

C1 = � � –1.43<br />

= 3.3<br />

From Equation 7-19 the correction factor C2 is<br />

C2 = 3.27(�P) –0.28 = 3.27 (50) –0.28 53 – 30<br />

�<br />

53<br />

= 1.0935<br />

TABLE 7-5 Typical Range of Sizes for Cyclones O<strong>pe</strong>rating at Pressures from 20 to 500<br />

kPa (3–72 psi)<br />

Diameter Diameter<br />

(of swirling (of swirling<br />

chamber) chamber)<br />

in mm Capacity in L/s in inches Capacity in USgpm<br />

100 1 L/s @ 20 kPa–6 L/s @500 kPa 4 16 gpm@ 3psi–96 gpm @ 72 psi<br />

150 3 L/s @ 20 kPa–15 L/s @500 kPa 6 48 gpm @ 3psi–240 gpm @ 72 psi<br />

250 7 L/s @ 20 kPa–35 L/s @500 kPa 10 110 gpm@ 3psi–555 gpm @ 72 psi<br />

380 12 L/s @ 20 kPa–60 L/s @500 kPa 15 190 gpm@ 3psi–950 gpm @ 72 psi<br />

510 26 L/s @ 20 kPa–140 L/s @500 kPa 20 410 gpm@ 3psi–2200 gpm @ 72 psi<br />

660 50 L/s @ 20 kPa–250 L/s @500 kPa 26 793 gpm@ 3psi–3963 gpm @ 72 psi<br />

760 85 L/s @ 20 kPa–450 L/s @500 kPa 30 1350 gpm@ 3psi–7100 gpm @ 72 psi


7.38 CHAPTER SEVEN<br />

TABLE 7-6 Typical Cyclone Size versus Particle Cut<br />

Diameter of hydrocyclone Discharge cut size<br />

The final correction factor C3 is based on the density of the solids with res<strong>pe</strong>ct to the<br />

liquid density:<br />

1650<br />

C3 = �� = 0.66<br />

�� 4800 – 1000<br />

d50C application = 23.77 × 3.3 × 1.0935 × 0.66 = 57 �m<br />

The flow rate in the underflow is determined from nozzle equations:<br />

Q = C D · A�2���P�/�� = 0.5 · 0.00503 �2�0� = 0.01124 m 3 /s = 11.24 L/s = 178 USgpm<br />

7-7-4 Magnetic Separators<br />

In beach and mineral sand plants as well as in taconite processing plants, minerals have<br />

magnetic pro<strong>pe</strong>rties. The presence of a magnet would attract the ferrous ores and separate<br />

them from other solids. This is the principle of magnetic separation. Magnetic separators<br />

work on two principles.<br />

1. An electromagnetic drum set in a stream<br />

2. A belt driven by an electromagnetic drum on which solids in a dry state or <strong>slurry</strong> form<br />

are allowed to pass to separate the ferrous ores<br />

7-8 FLOTATION CIRCUITS<br />

25 mm (1�) 5 �m (mesh 2500)<br />

100 mm (4�) 40 �m (mesh 380)<br />

250 mm (10�) 75 �m (mesh 200)<br />

500 mm (20�) 150 �m (mesh 100)<br />

Flotation is a method of separating solids from streams by creating a froth to which they<br />

are attracted. Thus in a <strong>slurry</strong> circuit, flocculants are added to create a froth rich with the<br />

metal concentrate. The trick is to make mineral particles hydrophobic, or water re<strong>pe</strong>llant.<br />

Flotation involves the selected “adsorption” of hydrocarbons (e.g., ethyl xanthate) on liberated<br />

minerals (e.g., chalcopyrite), which can then be attached to and transported by air<br />

bubbles in the <strong>slurry</strong> to a so-called froth layer and then separated from the hydrophilic<br />

(wetted) particles.<br />

For flotation to be efficient, it must be re<strong>pe</strong>ated a few times in a circuit that includes a<br />

rougher, a scavenger, and a cleaner as shown in Figure 7-28.<br />

The collector in a flotation circuit consists of a hydrophobic hydrocarbon chain of<br />

melecules (grease or wax) that re<strong>pe</strong>ls the mineral-laden water and causes it to attach itself<br />

to the passing air bubble. The surface chemistry is divided into three categories:


feed<br />

circulating<br />

tails<br />

1. Physical adsorption with a free energy of the collector smaller than 5 kcal/mol<br />

2. Chemisorption with a free energy of the collector larger than 30 kcal/mol<br />

3. Intermediary stages between adsorption and chemisorption<br />

Sulfide minerals are relatively easier to separate by chemisorption because they can<br />

use the major collectors such as xanthates and dithiosphophates. Certain s<strong>pe</strong>cial additives<br />

with high surface energy capabilities can also be added to separate different grades of sulfides<br />

(e.g., to sink pyrite while floating chalcopyrite).<br />

Oxide minerals (e.g., hematite, apatite. etc.) and silicate minerals are more difficult to<br />

separate by flotation than sulfide minerals. For oxides and silicate minerals, flotation is<br />

difficult because it is done by adsorption with minimal free surface energy using anionic<br />

fatty acids and cationic amines, which o<strong>pe</strong>rate essentially by electrostatic forces.<br />

When various ores are present, flotation may be done in stages using tanks in series. In<br />

each tank, a different pH level may be set or different collectors may be added, with the<br />

output from each tank going to a different circuit for further treatment.<br />

Depressants are chemicals that make the particle surface hydrophilic and nonfloatable.<br />

Typical depressants include bichromate, cyanide, zinc sulphate, and lime.<br />

Activators are chemicals that make the surface of nonfloating particles active for collector<br />

attachment. Typical activators are cop<strong>pe</strong>r sulphate and sodium sulphide.<br />

The pH value is a determining factor in many flotation circuits. It is adjusted by using<br />

various chemicals such as lime, caustic soda, sulfuric acid, etc.<br />

Frothers are chemicals that are used to decrease the surface tension of water in order to<br />

� Develop improved stability in the pulp<br />

� Achieve smaller and better bubble size<br />

COMPONENTS OF SLURRY PLANTS<br />

rougher<br />

cleaner<br />

Air bubble Water<br />

particles<br />

circulating concentrate<br />

concentrate<br />

concentrate<br />

tails scavenger tails<br />

FIGURE 7-28 Flow chart for flotation circuit with rougher, scavenger, and cleaner.<br />

Fig 7-28<br />

7.39


7.40 CHAPTER SEVEN<br />

� Create a suitable froth layer<br />

� Help destroy froth, after which it is removed<br />

Typical frothers include alcohols and pine oil.<br />

The size of the flotation tank is based on the required flow rate as well as the required<br />

retention time, an aeration factor, and a scale factor:<br />

Q·tr · sc volume = � (7-21)<br />

a<br />

where<br />

s c = scale factor = 0.85 for plants<br />

= 1.0 for pilot plants<br />

= 1.7 for lab batch<br />

Q = flow rate<br />

t r = retention time<br />

a = aeration factor = 0.85<br />

Example 7-3<br />

A flow rate of 500 m3 /hr requires a retention time of 20 minutes. Size the tank assuming<br />

four tanks.<br />

volume = (500/60) · 20 · 1/0.85 = 196 m3 196/4 = 49 m3 <strong>pe</strong>r tank<br />

The number of cells in a flotation circuit is determined by the degree of metallurgical<br />

control and the concern for short-circuiting. It used to be believed that the correct approach<br />

would consist of small cells and longer banks. However, with the advent of s<strong>pe</strong>cial<br />

mixers and good aeration techniques, it is now possible to use larger tanks (Figure 7-29).<br />

Flotation circuit can be very simple or very complex. A simple circuit such as used<br />

with coal, achieves floatation in a single step and does not involve cleaning of the froth.<br />

In a more complex circuit, an initial stage, called the rougher, is added; it acts as a preconcentrator.<br />

The flocculated output goes then to a second stage, called cleaning, that is<br />

done at higher dilution and is sometimes associated with regrinding at various stages.<br />

When the ore grade is fairly low but the mineral is of high value, a scavenger is used<br />

for additional preconcentration.<br />

Froth is a real challenge in the design of pumps. This will be reviewed in Chapter 8.<br />

7-9 MIXERS AND AGITATORS<br />

Mixers or agitators (Figure 7-30) are very important components of mineral and chemical<br />

process plants. They are used in various stages such as flotation circuits, leaching circuits,<br />

gold adsorption on carbon, preparation of s<strong>pe</strong>cial chemicals such as milk of lime, preparation<br />

of feed for pi<strong>pe</strong>lines, and final storage where sedimentation is likely to occur.<br />

Mixers are used in the gold leaching processes. S<strong>pe</strong>cial tanks for carbon in pulp (CIP)<br />

or carbon in leach (CIL) are built with mixers. The largest diameter of these tanks is approximately<br />

17 m (56 ft).<br />

For large plants, the process of flotation or leaching in a single tank is not very efficient.<br />

To increase productivity, tanks are installed in series (up to five stages or five tanks<br />

in a series), thus eliminating possible short-circuiting.


feed<br />

air<br />

agitator im<strong>pe</strong>ller<br />

COMPONENTS OF SLURRY PLANTS<br />

concentrate<br />

FIGURE 7-29 Simplified flotation circuit.<br />

gangue<br />

FIGURE 7-30 Top-entry agitator. (Courtesy of Hayward Gordon, Canada.)<br />

7.41


T<br />

H = D<br />

7.42 CHAPTER SEVEN<br />

There are processes that require a single mixing tank with one or two agitator mixers;<br />

an example is the preparation milk of lime, where solids are simply mixed with water and<br />

the mixers are used to prevent sedimentation.<br />

S<strong>pe</strong>cial processes in solvent extraction plants require gentle mixing while pumping the<br />

organic solution over a weir. For example, in a cop<strong>pe</strong>r SX-EW plant the organic solution<br />

is kerosene-based; it absorbs the cop<strong>pe</strong>r sulfates and separates them from other solids. For<br />

these applications, manufacturers have develo<strong>pe</strong>d s<strong>pe</strong>cial pump mixers, which are essentially<br />

large, o<strong>pe</strong>n shrouded pump im<strong>pe</strong>llers with a large number of vanes.<br />

Most tanks used in mining are built with a height to diameter ratio around unity and use<br />

a single-stage mixer. The mixer is of a vertical shaft design (Figure 7-30). The im<strong>pe</strong>ller diameter<br />

is in the range of 30% to 45% of the tank diameter. The im<strong>pe</strong>ller is usually situated<br />

at about 30% of the depth of the tank. Baffles at the wall of the tank break the vortex that is<br />

formed by the agitator. These baffles have a width of 8% of the tank diameter.<br />

Certain processes use tall, concentric tanks (sometimes called Pechuka tanks) with<br />

two agitators in a series on a single shaft. These are more common in South Africa than in<br />

North America.<br />

Horizontal agitators are installed on the <strong>pe</strong>riphery of very large tanks, particularly in<br />

the pulp and pa<strong>pe</strong>r industry. They have not been popular in <strong>slurry</strong> mixing tanks in mineral<br />

processes as they are difficult to maintain.<br />

A vertical mixing tank (Figure 7-31) is fit with baffles at the walls to break the vortex<br />

generated by the agitator. A certain gap is left between the baffles and the wall of the<br />

tank.<br />

In some res<strong>pe</strong>cts, the pro<strong>pe</strong>ller-ty<strong>pe</strong> mixer causes continuous mass flow against a<br />

stationary flat bottom. If the s<strong>pe</strong>ed of rotation is not sufficient, a stagnation area develops.<br />

The levels of turbulence in the tank, as well as the sha<strong>pe</strong> of the bottom of the tank,<br />

feed<br />

D = 0.3 D<br />

A T<br />

to 0.45 D<br />

T<br />

D T<br />

b = D T/12 to D T/10<br />

gap =<br />

(1/72)<br />

tank<br />

diameter<br />

FIGURE 7-31 Typical dimensions for the design of mixing tanks.<br />

C = DA<br />

= DT<br />

/3


COMPONENTS OF SLURRY PLANTS<br />

7.43<br />

are important parameters. Sometimes a conical deflector is installed at the bottom of the<br />

tank.<br />

Mixing can be done by using an outside pump. The tank must then have a conical bottom.<br />

The discharge of the tank at the bottom flows to a pump. The discharge of the pump<br />

is returned to the tank and passes through a jet mixer.<br />

For tanks with a flat bottom, the discharge may be from the bottom or through a pi<strong>pe</strong><br />

on the side (Figure 7-32).<br />

For very viscous mixtures, anchor agitators are recommended. The blades are vertical<br />

and rotate fairly close to the wall surface of the tank. For some difficult and frothy pulps<br />

in biological <strong>slurry</strong> treatment, helicoidal mixers are installed (Figure 7-33).<br />

For some complex mixtures, the agitator may incorporate a hollow shaft to sparge<br />

oxygen, an im<strong>pe</strong>ller to break up the froth at a high level, and one at the bottom of the shaft<br />

to mix the <strong>slurry</strong> (Figure 7-34).<br />

The pro<strong>pe</strong>ller-ty<strong>pe</strong> agitator is the most common in the mining industry. Its design can<br />

be examined from various angles: mechanical strength, s<strong>pe</strong>ed of o<strong>pe</strong>ration, hydrofoil<br />

sha<strong>pe</strong> of the blades, etc. The shaft is designed for the “jamming” condition or “start-up”<br />

in a settled tank. The main force is taken at 75% of the maximum radius blade span meas-<br />

feed feed<br />

bottom and side discharge bottom and central discharge<br />

feed feed<br />

Side discharge<br />

Top discharge<br />

FIGURE 7-32 Various patterns of discharge from the mixing tank in accordance with the required<br />

degree of agitation.


Anchor mixer for very viscous mixtures Helicoidal mixer for very Helicoidal viscous mixer mixtures for v<br />

FIGURE 7-33 Anchor and helicoidal mixers are used for particularly viscous mixtures.<br />

So Sole lepl plate ate to suppor support t<br />

Top of vessel vessel structure<br />

mixer<br />

Top column<br />

(flanged)<br />

foam breaker<br />

Bottom Bo ttom column (flanged)<br />

oxygen discharge<br />

7.44<br />

hollow shaft vertical motor<br />

Ho Hollow llow shaft for for<br />

central oxygen oxygen flow<br />

threaded coupling<br />

Op O<strong>pe</strong>n en housing for for<br />

foam breaker<br />

Gre Grease ase lubricated sleeve<br />

bearing & packing packing<br />

Mixer<br />

FIGURE 7-34 Complex mixer for biological slurries with central injection of oxygen and<br />

with a baffle to skim the froth.


COMPONENTS OF SLURRY PLANTS<br />

7.45<br />

ured from the center of the shaft. For severe duty application, the shaft is designed for 2.5<br />

times the rated motor torque without exceeding the yield stress (or 0.2% proof stress) of<br />

the shaft material. For light-duty mixers, the shaft is designed for 1.5 times the rated motor<br />

torque without exceeding the yield stress of the shaft material.<br />

The shaft must not o<strong>pe</strong>rate at s<strong>pe</strong>eds higher then 70% of the first critical s<strong>pe</strong>ed if the<br />

im<strong>pe</strong>ller is not dynamically balanced, or higher than 85% of the first critical s<strong>pe</strong>ed if the<br />

im<strong>pe</strong>ller is dynamically balanced. The deflection of the shaft should be limited, particularly<br />

in the case of anchor agitators, so that the blades do not hit the walls or baffles.<br />

The shaft is supported between two bearings in a cantilever arrangement. The bearings<br />

may be part of a vertical gearbox or an inde<strong>pe</strong>ndent bearing assembly. For large agitators,<br />

the gearbox and bearing assembly are integral.<br />

In Figure 7-35, the flow between the two levels z1 and z2 contracts across the pro<strong>pe</strong>ller,<br />

which induces a velocity Vi. Since the flow at the free surface as well as at the bottom<br />

of the tank is negligible, this flow resembles the ground effect of a hovering helicopter.<br />

From airscrew theory, the thrust across the pro<strong>pe</strong>ller is:<br />

2 T = 2A�V i (7-22)<br />

where<br />

Vi = the induced velocity<br />

A = area of flow across the pro<strong>pe</strong>ller<br />

Vi = �T�/2�A��� (7-23)<br />

where<br />

T = thrust<br />

3 power = TVi = 2�AVi (7-24)<br />

Another important theory used to calculate thrust is the blade theory. In Chapter 3,<br />

the concept of lift and drag around an aerofoil or an aircraft wing was introduced. The<br />

blade of a pro<strong>pe</strong>ller mixer is essentially a rotating wing exposed to a flow velocity V<br />

and a rotating s<strong>pe</strong>ed in rpm. Due to the contraction of the stream across the pro<strong>pe</strong>ller,<br />

the flow velocity is half the induced velocity. Since the blades are set at a certain pitch<br />

angle � (quite often 40–45°), they are at an angle of attack with res<strong>pe</strong>ct to the relative<br />

s<strong>pe</strong>ed. The relative s<strong>pe</strong>ed is the vectorial addition of these two <strong>pe</strong>r<strong>pe</strong>ndicular s<strong>pe</strong>eds<br />

(Figure 7-35).<br />

2 2 2 W= �¼�V� i�+� r� �� � (7-25)<br />

where the angular s<strong>pe</strong>ed � = 2�N/60 and N is rotations <strong>pe</strong>r minute.<br />

When the flow approaches a blade at a relative s<strong>pe</strong>ed W and an angle of incidence �<br />

with res<strong>pe</strong>ct to the chord of the blade, a certain pressure distribution develops around the<br />

blade. The result is a lift force L <strong>pe</strong>r<strong>pe</strong>ndicular to W and a resistance drag force D tangential<br />

to it. A good designer keeps the angle of attack at a value that corresponds to the maximum<br />

lift-to-drag ratio. For every airfoil, a plot of lift-to-drag curve (Figure 7-36) is obtained.<br />

If the blade is pitched at an angle �, then the vertical force is<br />

Y = lift cos � – drag sin � = L cos � – D sin � (7-26)<br />

and the horizontal force<br />

X = L sin � + D cos � (7-27)<br />

The vertical force Y is opposite to thrust, whereas the horizontal force X multiplied by<br />

the radius gives the resistant torque.<br />

Near the tips of the pro<strong>pe</strong>ller, the flow degrades due to the presence of tip vortices.


7.46<br />

z<br />

1<br />

z 2<br />

feed<br />

V /2<br />

i<br />

V<br />

i<br />

b = D T/12 to D T/10<br />

gap<br />

angle of<br />

incidence<br />

W<br />

W<br />

V /2<br />

i<br />

Y = L cos � – D sin �<br />

Lift<br />

U = r<br />

pitch<br />

angle of<br />

blade<br />

Drag<br />

FIGURE 7-35 Induced velocity and hydrodynamic forces for a pro<strong>pe</strong>ller-ty<strong>pe</strong> mixer.<br />

W<br />

X = L sin � + D cos �


Lift coefficient<br />

C L<br />

1.3<br />

angle of<br />

incidence<br />

10 o<br />

COMPONENTS OF SLURRY PLANTS<br />

W<br />

stall<br />

Angle of attack<br />

(or incidence )<br />

Lift<br />

7.47<br />

The load distribution for many pro<strong>pe</strong>llers is a maximum at 75% of the radius (Figure 7-<br />

37) so that the effective torque is measured at this region.<br />

In order to predict the <strong>pe</strong>rformance of a full-scale mixer, a test may be conducted on a<br />

reduced scale model under laboratory conditions. Performance is scaled up to large units<br />

using nondimensional factors such as the power factor from the theory of rotating equipment.<br />

In basic terms, it means that two mixers of the same geometrical design (but differ-<br />

Drag<br />

Drag coefficient<br />

C<br />

D<br />

0.1<br />

stall<br />

10<br />

Angle of attack<br />

(or incidence )<br />

o<br />

FIGURE 7-36 Lift and drag forces as a function of the angle of incidence with res<strong>pe</strong>ct to the<br />

flow.<br />

radial distribution of load mixer pro<strong>pe</strong>ller<br />

blade<br />

FIGURE 7-37 Distribution of the total force as a function of the span of the blade.


7.48 CHAPTER SEVEN<br />

ent sizes) will behave in similar ways with the same fluid and tank conditions. They<br />

would have the same power number. The power number is defined as<br />

P<br />

Cp = � (7-28)<br />

3 5 N D �<br />

where<br />

D = tip diameter of the mixer<br />

N = rotational s<strong>pe</strong>ed (rpm)<br />

� = density of the <strong>slurry</strong><br />

P = power<br />

The mixer Reynolds number is defined as<br />

ND<br />

Rem = (7-29)<br />

2� �<br />

2�<br />

The relationship between the power number and the Reynolds number is shown in<br />

Figure 7-38. Examples of the power number are presented in table Table 7-7.<br />

The ability of the mixer im<strong>pe</strong>ller to pump or induce flow is measured and defined by a<br />

nondimensional flow factor:<br />

Q<br />

CQ = � (7-30)<br />

3 ND<br />

It is also a function of the Reynolds number, as shown in Figure 7-39.<br />

Gates et al. (1976) examined the use of mixerss to maintain solids in sus<strong>pe</strong>nsion. An<br />

equivalent volume Vol eq is defined is defined as<br />

FIGURE 7-38 Power coefficient versus Reynolds number. The top curve is typical of flatblade<br />

mixers with wide blades. The middle curve is typical of flat-blade mixers with narrow<br />

blades. The bottom curve is typical of pitched-blade mixers. (Reproduced by <strong>pe</strong>rmission of<br />

Hayward Gordon.)


COMPONENTS OF SLURRY PLANTS<br />

TABLE 7-7 Typical Power Number for Mixers in Turbulent Flows<br />

Ty<strong>pe</strong> 3 Blades 4 Blades<br />

45° flat-pitched blades 1.6 1.7<br />

Pro<strong>pe</strong>ller (marine) 0.7 0.8<br />

Hydrofoil 0.2 0.5<br />

7.49<br />

Voleq = Sm Vol (7-31)<br />

where Sm is the s<strong>pe</strong>cific gravity of the <strong>slurry</strong> mixture.<br />

The terminal velocity for spheres was discussed in Section 3-1-3-1. A design settling<br />

velocity Vd is correlated to Vt by a correction factor fw: Vd = fwVt (7-32)<br />

where<br />

Vd = design settling velocity<br />

Vt = the terminal (or free settling) s<strong>pe</strong>ed<br />

The correction factor fw is presented in Table 7-8 as a function of weight concentration.<br />

This empirical coefficient was develo<strong>pe</strong>d by Chemineer Inc., based on ex<strong>pe</strong>rimental<br />

work. It is often difficult to predict the nature of the flow or the drag coefficient near an<br />

im<strong>pe</strong>ller blade. At weight concentrations in excess of 15%, the solids start to interact, hindering<br />

settling so that the settling velocity must be adjusted.<br />

The level of agitation is very important to the mechanics of sus<strong>pe</strong>nsion. Chemineer<br />

Inc. develo<strong>pe</strong>d a scale of agitation from 1 to 10, summarized by Gates and al. (1976) as in<br />

Table 7-9. Figure 7-40 shows the level of sus<strong>pe</strong>nsion of solids in correlation with the<br />

Chemineer scale. Often, manufactures define mixing as simple, mild, medium, vigorous,<br />

or violent (Figure 7-40).<br />

The science of mixing and keeping solids in sus<strong>pe</strong>nsion is highly empirical. The engineer<br />

should take into account existing similar installations as well as lab work results.<br />

Because of wear associated with slurries, a simple flat blade system is used to design<br />

C =<br />

Q<br />

Flow Coefficient<br />

Q<br />

N D 3<br />

Laminar Transition<br />

Reynolds Number<br />

Turbulent<br />

2<br />

Re = ND /(2 )<br />

FIGURE 7-39 Flow coefficient versus Reynolds number.<br />

D/H = 0.40<br />

D/H = 0.45<br />

D/H = 0.50<br />

Ratio of Im<strong>pe</strong>ller<br />

diameter to tank height


7.50 CHAPTER SEVEN<br />

TABLE 7-8 The Correction Factor f w Presented as a Function of<br />

Weight Concentration<br />

Solids weight concentration (%) Factor f w<br />

2 0.8<br />

5 0.84<br />

10 0.91<br />

15 1.0<br />

20 1.1<br />

25 1.2<br />

30 1.3<br />

35 1.42<br />

40 1.55<br />

45 1.70<br />

50 1.85<br />

From Gates et al., 1976, reprinted by <strong>pe</strong>rmission from Chemical Engineering.<br />

TABLE 7-9 Chemineer Scale for Agitation of Solids in Sus<strong>pe</strong>nsion<br />

Scale of agitation Description<br />

1–2 At levels 1–2, agitation is required for minimal sus<strong>pe</strong>nsion of solids.<br />

Agitators capable of working at an agitation level of 1–2 will:<br />

� Produce motion of all of the solids of the design-settling velocity in the<br />

vessel<br />

� Permit moving fillets of solids on the bottom, which are <strong>pe</strong>riodically<br />

sus<strong>pe</strong>nded<br />

3–5 Agitation levels 3–5 characterize most chemical process industries solids<br />

sus<strong>pe</strong>nsion applications. This scale range is typically used for dissolving<br />

solids.<br />

Agitators capable of working at an agitation level of 3–5 will:<br />

� Sus<strong>pe</strong>nd all of the solids of design velocity completely off the vessel<br />

bottom<br />

� Provide <strong>slurry</strong> uniformity to at least one-third of the fluid batch height<br />

� Be suitable for <strong>slurry</strong> draw-off at low exit-nozzle elevations<br />

6–8 Agitation levels 6–8 characterize applications where the solids sus<strong>pe</strong>nsion<br />

level approaches uniformity.<br />

Agitators capable of scale level 6 will:<br />

� Provide concentration uniformity of solids to 95% of the fluid batch<br />

height<br />

� Be suitable for <strong>slurry</strong> draw-off up to 80% of the fluid batch height<br />

9–10 Agitation levels 9–10 characterize applications where the solids sus<strong>pe</strong>nsion<br />

uniformity is the maximum practical.<br />

Agitators capable of scale 9 will:<br />

� Provide concentration uniformity of solids to 98% of the fluid batch<br />

height<br />

� Be suitable for <strong>slurry</strong> draw-off by means of overflow<br />

From Gates et al., 1976. Reprinted by <strong>pe</strong>rmission from Chemical Engineering.


a. Unstable particles are on<br />

vessel bottom (Scale of<br />

agitation = 1)<br />

the mixer. The blades are often pitched at an angle of 45° with res<strong>pe</strong>ct to the horizontal<br />

plane. To size such a flat-bladed pro<strong>pe</strong>ller in a mixing tank, with baffles at 90° to each<br />

other, baffle width of 1/12th of the tank diameter, and offset from the wall with a gap of<br />

1/72nd of the tank diameter, Gates et al. (1976) proposed the following empirical equation<br />

in USCS units:<br />

Din = 394� � 0.2 HP<br />

� (7-33a)<br />

SmN 3n where<br />

Din = diameter in inches<br />

HP = power in hp<br />

n = number of im<strong>pe</strong>llers<br />

N = s<strong>pe</strong>ed in rev/min<br />

Sm = s<strong>pe</strong>cific gravity of the <strong>slurry</strong> mixture<br />

To express Equation 7-33a in SI units:<br />

COMPONENTS OF SLURRY PLANTS<br />

b. Particles swept off vessel<br />

bottom (Scale of agitation = 3)<br />

Dimp = 37.57� � 0.2 P<br />

�<br />

SmN3n 7.51<br />

c. Solids are homogeneously<br />

distributed (Scale of<br />

agitation = 9)<br />

FIGURE 7-40 Intensity of agitation yields different patterns of solid sus<strong>pe</strong>nsion. (From<br />

Gates et al., 1976. Reproduced by <strong>pe</strong>rmission from Chemical Engineering.)<br />

(7-33b)<br />

where<br />

Dimp = diameter in meters<br />

P = power in Watts<br />

To correlate between the levels of agitation in a tank, the s<strong>pe</strong>ed of rotation, and diameter<br />

of the im<strong>pe</strong>ller, Gates et al. (1976) plotted the scale of agitation versus �, a factor defined<br />

in U.S. units as:<br />

� = N3.75 2.81 Dimp /Vd<br />

with Vd expressed in ft/min, D in inches, and N in rev/min (Figure 7-41).


7.52 CHAPTER SEVEN<br />

FIGURE 7-41 Chart to determine s<strong>pe</strong>ed and diameter of a mixer versus the solids sus<strong>pe</strong>nsion<br />

scale. [From Gates et al. (1976). Reprinted by <strong>pe</strong>rmission of Chemical Engineering.]<br />

Example 7-4<br />

A tank contains 600 ft3 of a <strong>slurry</strong> mixture. The s<strong>pe</strong>cific gravity of the solids is 4.1 and the<br />

average particle size d50 is 0.01 ft. It is required to be designed for overflow output at an agitation<br />

scale of 9. The weight concentration of the mixture is 20%. Size the mixer, assuming<br />

a single im<strong>pe</strong>ller with an im<strong>pe</strong>ller-to-tank diameter of 0.4, baffles 1/12th the tank diameter<br />

and baffle gap of 1/72nd the tank diameter; use Figure 7-34. Assume viscosity of 1<br />

cP. Compare the power consumption with a mixer o<strong>pe</strong>rating at a Chemineer scale of 5.<br />

Solution in USCS Units<br />

Since most tank have a diameter equal to the height, and assuming a volume of 90% the<br />

tank volume, the effective volume occupied by the <strong>slurry</strong> is<br />

3 3<br />

Vol = 0.9 × 0.25 × �DT = 600 ft<br />

Hence,<br />

DT = 9.47 ft<br />

T = 9.47/0.9 = 10.52<br />

Dimp = 0.4 × 10.52� = 4.21� or approximately 50.5 inch<br />

For a scale of 9 and D/T = 0.4, � = 30 × 10 10 , so<br />

� = N 3.75 D 2.81 /V d<br />

We must determine V d. The particles are coarse enough to assume a drag coefficient C D =<br />

0.44. Substituting into Equation 3-7<br />

4(�S – � 4(3.1) × 32.2 × 0.01<br />

CD =<br />

L)gdg �� = ��� 2 3 × V t<br />

3�LV t 2<br />

Vt = 1.5 ft/s = 90 ft/min<br />

From Table 7-8, the correction factor fw = 1.0 and Vd = VT, or<br />

� =30 × 10 10 = N3.75D2.81 /Vd = N3.7550.52.81 /90 = 679.5 N3.75 N = 202 rev/min<br />

To determine the required horsepower, Equation 7-33a is used:<br />

Din = 394� � 0.2 HP<br />

�<br />

SmN 3n


The s<strong>pe</strong>cific gravity of the mixture is obtained by using Equation 1-4:<br />

or S m = 1.128.<br />

COMPONENTS OF SLURRY PLANTS<br />

100<br />

���<br />

15/4100 + (100 – 15)/1000<br />

� m = = 1128 kg/m 3<br />

[50.5/394] × (1.128 × 202 3 ) 0.2 = HP 0.2<br />

or HP = 322 hp. For a scale of 5, and D/T = 0.4, � = 5 × 10 10 , so<br />

� = 5 × 10 10 = N 3.75 D 2.81 /V d = N 3.75 50.5 2.81 /90 = 679.5 N 3.75<br />

N = 125 rev/min<br />

To determine the required horsepower, Equation 7-33a is used:<br />

[50.5/394] × (1.128 × 1253 ) 0.2 = HP0.2 7.53<br />

or HP = 76 hp. Going from a mild level of agitation at a 5 on the Chemineer scale to level<br />

9 increases the power consumption 4.24 fold.<br />

The shear rate was previously defined in Chapter 2 as the rate of change of the velocity<br />

with res<strong>pe</strong>ct to height above the wall. More generally for a mixer, the shear rate is the<br />

change of velocity with res<strong>pe</strong>ct to depth. The induced velocity is essentially created by<br />

the im<strong>pe</strong>ller, and the maximum shear rate is ex<strong>pe</strong>rienced at the tip of the blade. There is,<br />

however, an average shear rate estimated for the im<strong>pe</strong>ller zone, and an average in bulk of<br />

the tank.<br />

For a radial im<strong>pe</strong>ller, all the flow is discharged at the tip of the im<strong>pe</strong>ller, and all the<br />

solids are subject to the same s<strong>pe</strong>ed and head. In the case of the pro<strong>pe</strong>ller or axial turbine,<br />

the velocity distribution is proportional to the radius. All solids pass through a higher<br />

shear zone in the case of the radial machine.<br />

The closer the im<strong>pe</strong>ller is to the bottom of the tank, the more the induced velocity is<br />

suppressed. A proximity factor h c/D is defined as the ratio of the gap of the im<strong>pe</strong>ller to the<br />

diameter. This principle is similar to those in the world of aeronautics. The ground effect<br />

is essentially the pressure exerted by a helicopter or airplane. The closer it is to the<br />

ground, the more pronounced is the effect and, eventually, the induced drag is reduced<br />

when the machine flies very close to the ground. In the case of the mixers, the flow is restricted<br />

as the im<strong>pe</strong>ller gets closer to the bottom of the tank. This affects power consumption.<br />

A radial mixer tends to induce flow tangentially, whereas an axial machine tends to<br />

induce it vertically. Pro<strong>pe</strong>ller and radial mixers tend to create different patterns of recirculation<br />

around their blades.<br />

Mounting more than one im<strong>pe</strong>ller on the same shaft gives different patterns of <strong>pe</strong>rformance<br />

based on the mutual interference that the different in<strong>pe</strong>llers exert on each other.<br />

The power from two axial turbines is not twice the power from a single pro<strong>pe</strong>ller, as the<br />

induced velocity of the second pro<strong>pe</strong>ller is not twice the induced velocity of the first pro<strong>pe</strong>ller.<br />

In fact, in some applications, the first pro<strong>pe</strong>ller is smaller in diameter than the bottom<br />

pro<strong>pe</strong>ller. However, two radial im<strong>pe</strong>llers closely spaced consume more than twice<br />

the power of a single unit.<br />

In sewage treatment plant processes as well as in chemical and mining processes, gas<br />

is sparged in the liquid. The im<strong>pe</strong>ller of the mixer is then used to provide dis<strong>pe</strong>rsion of the<br />

gas and circulation of the tank contents. For radial machines, a small radial im<strong>pe</strong>ller is installed<br />

at the bottom to mix the gas. In the case of hydrofoils, different approaches are<br />

used. If the shaft and blades are hollow, gas may be pum<strong>pe</strong>d through the shaft and blades.<br />

In other applications, the im<strong>pe</strong>ller is contained within a cylinder that is within the tank.<br />

This prevents flooding the im<strong>pe</strong>ller with gas.


7.54 CHAPTER SEVEN<br />

A hydrofoil ty<strong>pe</strong> of mixer would have the highest pumping capacity but would develop<br />

the lowest head or shear at constant horsepower or torque. Hydrofoil mixers are often<br />

chosen for high-flow applications.<br />

For heat transfer, solid sus<strong>pe</strong>nsion, blending, or solid dissolving, bulk pumping is critical.<br />

A hydrofoil mixer would be the preferred machine. However, for gas–liquid contacting,<br />

molecular mixing, solid dis<strong>pe</strong>rsion (reduction of agglomerates), an axial flow, flatblade<br />

machine or radial mixer are preferred options.<br />

The mixing of non-Newtonian slurries is fairly complex and must rely on ex<strong>pe</strong>rimental<br />

data. Wasp et al. (1977) proposed that the power consumption for mixing non-Newtonian<br />

slurries is a function of the Reynolds number, the Hedstrom number, and the<br />

Froude number:<br />

N<br />

= fn� , , � (7-34)<br />

In other words, the power consumption is based on the Reynolds number, Hedstrom number,<br />

and Froude number based on the im<strong>pe</strong>ller diameter.<br />

The mixing of non-Newtonian fluids is common in the manufacture of polymers and<br />

considerable data is available. It is, however, not very wise to extrapolate to non-Newtonian<br />

<strong>slurry</strong> mixtures.<br />

In the last 25 years of the 20th century, larger and larger mixers have been built. Certain<br />

shaft failures and blade failures have occurred on some large agitators. Some were<br />

due to corrosion of the bolts holding the blades and others due to jamming of the shaft. It<br />

is important to understand that starting an agitator from fully settled conditions can be<br />

very stressful to the machine.<br />

The reader should consult appropriate reference books on machine design and gearbox<br />

design. A service factor of 1.5 is a minimum for sizing the gearbox. It would be beyond<br />

the sco<strong>pe</strong> of this book to explore the selection of gearboxes for agitators. However, the<br />

designer of <strong>slurry</strong> mixers should be aware of certain important mechanical criteria, such<br />

as critical s<strong>pe</strong>ed.<br />

To examine the distribution of loads on the blade, the program “Agitblade” may be<br />

used.<br />

2 P<br />

2� �0 Dimp � � � � 3 5 2 2 2 �N D imp �NDimp �N Dimp g<br />

Program “Agitblade” for Pro<strong>pe</strong>ller Blade Loads<br />

CLS<br />

REM pro<strong>pe</strong>ller design for AGITATOR<br />

PI = 3.1415<br />

DIM R(20), PITCH(20), V(20), W(20), ALPHA(20), BETA(20), BLPHA(20)<br />

DIM C(20), CL(20), CD(20), CT(20), CU(20), T(20), INCIA(20)<br />

DIM P(20), TK(20), VV(20), VH(20), VS(20), PK(20)<br />

INPUT “FLUID DENSITY”; DENS<br />

INPUT “radius at hub “; RHUB<br />

INPUT “ RADIUS AT TIP”; RTIP<br />

AP = PI * RTIP ^ 2<br />

100 PRINT “OPTIONS FOR COMPUTATION”<br />

PRINT “1-CALCULATIONS ASSUMING KNOWLEDGE OF INDUCED VELOCITY”<br />

PRINT “ USING MOMENTUM THEORY”<br />

PRINT “2- BLADE THEORY CALCULATIONS FOR BLADEWISE DISTRIBUTION OF<br />

FORCES”<br />

PRINT “ AND FLOW CHARACTERISTICS”<br />

PRINT “3- VORTEX FLOW CALCULATIONS”


COMPONENTS OF SLURRY PLANTS<br />

INPUT “YOUR OPTION “; OPT<br />

IF OPT = 1 THEN 300<br />

IF OPT = 2 THEN 2000<br />

IF OPT = 3 THEN 6000<br />

300 INPUT “NUMBER OF PROPELLERS ON THE SAME SHAFT”; NU<br />

IF NU > 1 THEN PRINT “CALCULATIONS FOR FIRST PROPELLER”<br />

INPUT “VELOCITY UPSTREAM THE PROPELLER “; VS<br />

IF VS = 0 THEN 305<br />

INPUT “IS THIS VELOCITY PARALLEL TO SHAFT (Y/N)”; V$<br />

IF V$ = “Y” OR V$ = “y” THEN 305<br />

INPUT “INCLINATION OF U/STREAM VELOCITY WRT SHAFT AXIS IN DEGREES”;<br />

INC<br />

INCI = INC * PI/180<br />

305 FOR M = 1 TO NU<br />

VS(M) = VS<br />

PRINT<br />

PRINT<br />

PRINT “CALCULATION FOR PROPELLER “; M<br />

INCIA(M) = INCI * 180/PI<br />

VV(M) = VS * COS(INCI)<br />

VH(M) = VS * SIN(INCI)<br />

GOTO 320<br />

310 VV(M) = VS<br />

320 IF VV(M) = 0 THEN PRINT “COMPONENT OF U/S VEL PARALLEL TO PROP<br />

IS NIL”<br />

321 INPUT “IS THE HYDROSTATIC PRESSURE EQUAL ON BOTH SIDES OF THE<br />

PROP (Y/N)”; PH$<br />

IF PH$ = “Y” OR PH$ = “y” THEN 340<br />

INPUT “HYDROSTATIC PRES UPSTREAM PROP “; PHUS<br />

INPUT “HYDROSTATIC PRES DOWNSTREAM PROP “; PHDS<br />

340 INPUT “DO YOU KNOW THE INDUCED VELOCITY (Y/N)”; IV$<br />

IF IV$ = “N” OR IV$ = “n” THEN 600<br />

INPUT “INDUCED VELOCITY “; VU(M)<br />

VU = VU(M)<br />

T(M) = DENS * AP * SQR((VV(M) + VU) ^ 2 + VH(M) ^ 2) * VU ^ 2 +<br />

(PHUS – PHDS) * AP<br />

P(M) = T(M) * VV(M) + .5 * T(M) * VU(M)<br />

TK(M) = T(M)/1000<br />

PRINT USING “ THRUST = ###########.#### kN”; TK(M)<br />

PK(M) = P(M)/1000<br />

PRINT USING “POWER = #########.### kW”; PK(M)<br />

VH = VH(M)<br />

VS = SQR((VV(M) + VU) ^ 2 + VH(M) ^ 2)<br />

INCI = ATN(VH/(VV(M) + VU(M)))<br />

NEXT M<br />

LPRINT “CALCULATIONS BASED ON MOMENTUM THEORY”<br />

FOR M = 1 TO NU<br />

7.55


7.56 CHAPTER SEVEN<br />

LPRINT<br />

LPRINT<br />

LPRINT “CALCULATION FOR PROPELLER “; M<br />

LPRINT USING “VELOCITY UPSTREAM SHAFT = ###.## m/s”; VS(M)<br />

LPRINT USING “ITS COMPONENT PARRALLEL TO SHAFT= ###.## m/s “; VV(M)<br />

LPRINT USING “ITS COMPONENT PERPANDICULAR TO SHAFT = ###.## m/s “;<br />

VH(M)<br />

LPRINT USING “ITS INCLINATION WRT TO SHAFT = ###.# deg “; INCIA(M)<br />

LPRINT USING “INDUCED VELOCITY = ###.## m/s “; VU(M)<br />

LPRINT USING “ RESULTANT THRUST = #####.## KN”; TK(M)<br />

LPRINT USING “ INDUCED POWER CONSUMPTION = #####.## KW”; PK(M)<br />

NEXT M<br />

GOTO 8000<br />

600 PRINT “YOU MAY HAVE TO CALCULATE THE INDUCED VELOCITY BY<br />

ASSUMING A CERTAIN THRUST MAGNITUDE”<br />

2000 INPUT “VERTICAL SPEED”; V<br />

INPUT “PITCH AT HUB”; PITCHR<br />

INPUT “PITCH AT TIP”; PITCT<br />

INPUT “REQUIRED TIP SPEED “; VTIP<br />

RPS = VTIP/RTIP<br />

PRINT USING “THE ROTATIONAL SPEED IN ######.## RAD/S “; RPS<br />

RDIV = (RTIP - RHUB)/10<br />

PITDIV = (PITCHR - PITCT)/10<br />

INPUT “chord at the root”; CR<br />

INPUT “TIP CHORD “; CT<br />

REM CALCULATE THE ADVANCE RATIO OF THE PROPELLER<br />

J = V/(2 * PI * RTIP)<br />

PRINT “ADVANCE RATIO OF PROP “; J<br />

REM CALCULATE BLADE AREA AND HENCE ASPECT RATIO<br />

SB = (RTIP - RHUB) * .5 * (CT + CR)<br />

AR = (RTIP - RHUB) ^ 2/SB<br />

PRINT “BLADE ASPECT RATIO”; AR<br />

PRINT “BLADE AREA “; SB<br />

CD0 = .015<br />

K = .1/(1 + 2/AR)<br />

KD = K ^ 2/3 * AR<br />

REM K IS THE LIFT COEFFICIENT SLOPE DCL/DALPHA IN THE LINEAR RANGE<br />

PRINT “LIFT SLOPE PER DEGREE IN THE LINEAR RANGE “; K<br />

ACR = (CR - CT)/(RTIP - RHUB)<br />

CDIV = (CR - CT)/10<br />

DTIP = 2 * RTIP<br />

INPUT “POWER NUMBER “; CP<br />

P = CP * DENS * RPS ^ 3 * DTIP ^ 5 * CP/2 * PI<br />

PK = P/1000<br />

TOR = PK/RPS<br />

PRINT USING “REQUIRED TORQUE = #####.## KNm”; TOR<br />

PRINT USING “REQUIRED POWER #####.## KW”; PK<br />

FOR N = 1 TO 11<br />

R(N) = RHUB + (N - 1) * RDIV<br />

PITCH(N) = PITCHR - (N - 1) * PITDIV


COMPONENTS OF SLURRY PLANTS<br />

V(N) = R(N) * RPS<br />

W(N) = SQR(V(N) ^ 2 + V ^ 2)<br />

BLPHA(N) = ATN(V/V(N))<br />

ALPHA(N) = BLPHA(N) * 180/(2 * PI)<br />

BETA(N) = PITCH(N) - ALPHA(N)<br />

C(N) = CR - ACR * (R(N) - RHUB)<br />

CL(N) = K * BETA(N)<br />

CD(N) = CD0 + KD * CL(N)<br />

CT(N) = CL(N) * COS(BLPHA(N)) - CD(N) * SIN(BLPHA(N))<br />

CU(N) = CL(N) * SIN(BLPHA(N)) + CD(N) * COS(BLPHA(N))<br />

T(N) = CT(N) * .5 * W(N) ^ 2 * DENS * C(N) * RDIV<br />

NEXT N<br />

RPM = RPS * 60/(2 * PI)<br />

LPRINT “AGITATOR PROPELLER”<br />

LPRINT USING “VERTICAL SPEED = #####.### m/s”; V<br />

LPRINT USING “HUB RADIUS = ###.#### m TIP Radius = ####.#### m “;<br />

RHUB, RTIP<br />

LPRINT USING “TIP SPEED = ####.### M/S”; VTIP<br />

LPRINT USING “ANGULAR SPEED = ######.## RPM”; RPM<br />

LPRINT USING “VERTICAL SPEED = ####.## m/s “; V<br />

LPRINT USING “PITCH AT TIP= ####.## ; PITCH AT HUB = ####.##”;<br />

PITCT, PITCHR<br />

LPRINT “BLADE AREA”; SB<br />

LPRINT “BLADE ASPECT RATIO”; AR<br />

LPRINT “APPROX LIFT SLOPE COEF IN LINEAR RANGE”; K<br />

LPRINT “APPROX LIFT/DRAG POLAR RATIO “; KD<br />

LPRINT USING “PROPELLER AREA ######.## m^2”; AP<br />

LPRINT “LOCAL RADIUS PITCH ANGLE ANGLE OF INCIDENCE TOTAL VELOCITY<br />

ALPHA”<br />

FOR N = 1 TO 11<br />

LPRINT R(N), PITCH(N), BETA(N), W(N), ALPHA(N)<br />

NEXT N<br />

LPRINT “ LOCAL RADIUS,CHORD, LIFT COEF, DRAG COEF, THRUST COEF,<br />

TANG FORCE COEF”<br />

7.57<br />

FOR N = 1 TO 11<br />

LPRINT R(N), C(N), CL(N), CD(N), CT(N), CU(N)<br />

LPRINT<br />

NEXT N<br />

6000<br />

8000<br />

INPUT “DO YOU WANT TO PROCEED WITH OTHER THEORIES OF DESIGN (Y/N) “;<br />

D$<br />

IF D$ = “Y” OR D$ = “y” THEN 100<br />

END<br />

REM CT=THRUST COEFFICIENT PARALLEL TO PROPELLER SHAFT<br />

REM CU= FORCE COEFFICIENT PERPANDICULAR TO SHAFT AND CAUSING<br />

RESISTANCE


7.58 CHAPTER SEVEN<br />

The designer of mixers with hydrofoils should be aware that the center of pressure<br />

does not always correspond with the center of gravity of the blade. The center of pressure<br />

could be as far as 25% of the blade chord on high-as<strong>pe</strong>ct-ratio blades or those with a high<br />

ratio of blade length to blade chord. A pitching–bending moment results; it must be considered<br />

when sizing the bolts of the blade.<br />

It is important to check the stress on the shaft of an agitator. The following program,<br />

written in QuickBasic, allows one to check for the case of a shaft with two im<strong>pe</strong>llers in<br />

series.<br />

Program “Dblagit.Bas” for Two Agitators Mounted on a Shaft<br />

PRINT “double agit”<br />

pi = 4 * ATN(1)<br />

‘INPUT “are you using SI units (Y/N)”; l$<br />

l$ = “n”<br />

IF l$ = “n” OR l$ = “N” THEN conv = .0254<br />

IF l$ = “y” OR l$ = “Y” THEN conv = 1!<br />

‘INPUT “distance from bottom bearing to first coupling”; ab<br />

ab = 9.41<br />

‘INPUT “shaft O.D. for first section of shaft ab “; dab0<br />

‘INPUT “shaft I.D. for first section of shaft ab “; dabi<br />

dab0 = 5<br />

jab = (pi/32) * (dab0 ^ 4 - dabi ^ 4) * conv ^ 4<br />

‘INPUT “distance from first coupling to second coupling”; bc<br />

bc = 135.64<br />

‘INPUT “shaft O.D. for second section of shaft ab “; dbc0<br />

‘INPUT “shaft I.D. for second section of shaft ab “; dbci<br />

dbc0 = 8.625<br />

dbci = 7.625<br />

‘sch 80<br />

jbc = (pi/32) * (dbc0 ^ 4 - dbci ^ 4) * conv ^ 4<br />

‘INPUT “distance from second coupling to first im<strong>pe</strong>ller”; cd<br />

cd = 48<br />

‘INPUT “shaft O.D. for third section of shaft cd “; dcd0<br />

‘INPUT “shaft I.D. for first section of shaft cd “; dcdi<br />

‘dcd0 = 6.625<br />

‘dcdi = 6.065<br />

dcd0 = dbc0<br />

dcdi = dbci<br />

jcd = (pi/32) * (dcd0 ^ 4 – dcdi ^ 4) * conv ^ 4<br />

‘INPUT “distance from first to second im<strong>pe</strong>ller”; de<br />

de = 150<br />

dde0 = dcd0<br />

ddei = dcdi<br />

jed = jcd<br />

‘INPUT “ power on top im<strong>pe</strong>ller”; p1<br />

p1 = 21 * 746<br />

‘INPUT “ power on bottom im<strong>pe</strong>ller”; p2<br />

p2 = 35 * 746<br />

INPUT “rpm”; rpm


w = rpm * 2 * pi/60<br />

‘assume start up torque factor 250%<br />

ts = 2.5<br />

t1 = p1/w<br />

t2 = p2/w<br />

PRINT “torque from top im<strong>pe</strong>ller”; t1<br />

‘INPUT “diameter of first im<strong>pe</strong>ller”; dimp1<br />

dimp1 = 84<br />

‘INPUT “diameter of second im<strong>pe</strong>ller”; dimp2<br />

dimp2 = 66<br />

r1 = dimp1 * conv/2<br />

r2 = dimp2 * conv/2<br />

f1 = t1/(.75 * r1)<br />

PRINT “ tangential force on top im<strong>pe</strong>ller”; f1<br />

f2 = t2/(.75 * r2)<br />

PRINT “tangential force on bottom im<strong>pe</strong>ller”; f2<br />

ma = (f1 * conv * (ab + bc + cd) + f2 * conv * (ab + bc + cd +<br />

de))/10<br />

‘assume unsymmetry of forces at 10% radius<br />

may = SQR(ma ^ 2 + .75 * (t1 + t2) ^ 2)<br />

mayn = ma/1000000!<br />

PRINT USING “total bending moment at a = ####.## MNm”; mayn<br />

syab = mayn/jab<br />

PRINT USING “shaft stress syab = ###.## MPa”; syab<br />

IF syab > 100 THEN PRINT “warning the required stress limit is 100<br />

MPa”<br />

mb = (f1 * conv * (bc + cd) + f2 * conv * (bc + cd + de))/10<br />

mby = SQR(mb ^ 2 + .75 * (t1 + t2) ^ 2)<br />

mbyn = mby/1000000!<br />

PRINT USING “total bending moment at b = ####.## MPa”; mbyn<br />

sybc = mbyn/jbc<br />

PRINT USING “shaft stress syab = ###.## MPa”; sybc<br />

IF sybc > 100 THEN PRINT “warning the required stress limit is 100<br />

MPa”<br />

mcd = (f1 * conv * cd + f2 * conv * (cd + de))/10<br />

mcd = SQR(mcd ^ 2 + .75 * t2 ^ 2)<br />

mcdn = mcd/1000000!<br />

PRINT USING “total bending moment at c = ####.## MPa”; mcdn<br />

sycd = mcdn/jab<br />

PRINT USING “shaft stress sycd = ###.## MPa”; sycd<br />

IF sycd > 100 THEN PRINT “warning the required stress limit is 100<br />

MPa”<br />

7-10 SEDIMENTATION<br />

COMPONENTS OF SLURRY PLANTS<br />

7.59<br />

Sedimentation is a form of separation of solids from liquids by using gravity forces rather<br />

than electrostatic, chemical (flotation), or magnetic forces. Sedimentation may be<br />

achieved by gravity forces, using thickeners and clarifiers. On the other hand, it may be<br />

accomplished by centrifugal forces, as in centrifuges. In gold extraction circuits, an inter-


7.60 CHAPTER SEVEN<br />

mediary centrifuge is sometimes installed between the hydrocyclones and the ball mill<br />

feed box. Centrifuges are sometimes called concentrators because they <strong>pe</strong>rmit the extraction<br />

of some of the heavy metals by applying a very high centrifugal force such as 60<br />

times the acceleration due to gravity (60 g).<br />

7-10-1 Gravity Sedimentation<br />

Gravity sedimentation is classified as thickening or increasing the concentration of the<br />

feed stream, or clarification or the removal of solids from relatively dilute streams. The<br />

former is used to prepare the feed for tailings and concentrate pi<strong>pe</strong>line flow, or for the removal<br />

of tailings on trucks. The latter is more frequently used in sewage and waste treatment<br />

plants, where the volume of solids is considerably smaller than in tailings and concentrate<br />

flows.<br />

Considerable research on the use of flocculants in the last quarter of the twentieth century<br />

has lead to more concentrated sedimentation with less thickener. It would be beyond<br />

the sco<strong>pe</strong> of this book to discuss all these new flocculants.<br />

In simple terms, a clarifier or a thickener is essentially a sedimentation tank. To make<br />

the sedimentation uniform, a rake or arm rotates slowly but continuously. A relatively<br />

clear layer of liquid forms at the top and is withdrawn through an overflow box feeding a<br />

launder. The <strong>slurry</strong> in the thickener is denser at lower and lower layers. The bottom of the<br />

thickener forms a shallow cone with the center feeding into an underflow pi<strong>pe</strong> to a separate<br />

launder or pump.<br />

The actual feed to the thickener is through a launder to the center. A feed box leads the<br />

<strong>slurry</strong> to a depth lower than the relatively clear water. Some s<strong>pe</strong>cial processes use intermediary<br />

mixing chambers where flocculants are added to accelerate the precipitation.<br />

The tank itself may be shallow and called a shallow thickener, or deep and called a<br />

deep thickener. The decision to choose either is often based on various parameters such as<br />

the final weight concentration, the rate of sedimentation, the viscosity, the design of the<br />

rake, as well as other parameters. This is at the basis of the design of the thickener (Figure<br />

7-42).<br />

The actual process of sedimentation in a tube is based on the settling (or terminal)<br />

s<strong>pe</strong>ed that was discussed at great length in Chapter 3. It is also depicted in Figure 7-43.<br />

Initially, the <strong>slurry</strong> is uniformly mixed. Gradually, the solids sink, forming three layers of<br />

liquid: free of solids, a dilute mixture, and a relatively dense layer. Eventually, all the<br />

solids in the dilute layer sediment out, leaving only two layers, one of water and one of a<br />

dense mixture with solids at minimum void ratio. The use of certain chemicals can accelerate<br />

the sedimentation of solids.<br />

The correlation between the terminal velocity of a sphere V t and the sedimentation<br />

s<strong>pe</strong>ed V s is correlated to the void fraction � (Cheremisinoff, 1984) by the following equation:<br />

Vs = Vt� 2X(�) (7-35)<br />

Where X(�) is a function of the void ratio that must be determined by tests. The void ratio<br />

is<br />

Vol f<br />

� = ��<br />

(7-36)<br />

Volf + Volp where<br />

Volp = volume filled by the particles<br />

Volf = volume of liquid filling the space between the particles


7.61<br />

FIGURE 7-42 Schematics of a thickener used for sedimentation of solids.


7.62 CHAPTER SEVEN<br />

height of dense phase<br />

fig 743<br />

For thickened sludges with a void ratio smaller than 0.7, Cheremisinoff (1984) proposed<br />

the following correlation:<br />

�<br />

Vs = 0.123 Vt� � (7-37)<br />

3<br />

Spheres can actually compact in a very dense pattern to a minimum void ratio of<br />

0.215, but Cheremisinoff (1984) indicated that the average void ratio from thickeners was<br />

0.6. For nonspherical and coarse particles, the situation becomes more complex because<br />

of the sha<strong>pe</strong> factor (discussed in Chapter 3), and it is the norm to conduct sedimentation<br />

tests on samples of the <strong>slurry</strong> before designing the thickener.<br />

7-10-2 Centrifuges<br />

clear water boundary<br />

dense phase boundary<br />

FIGURE 7-43 Response of gravity sedimentation with time.<br />

� 1 – �<br />

time (minutes)<br />

Centrifuges use centrifugal force as a means to separate solids from liquids. Liquid is fed<br />

into the inlet and a rotating bowl is used to apply the centrifugal force, similar to a clothes<br />

drier that separates liquid from clothes by continuously rotating the clothes. Obviously,<br />

with <strong>slurry</strong>, it is more complex (Figure 7-44).<br />

The centrifugal force is defined as<br />

F = mR�2 (7-38)<br />

where<br />

� = 2�N/60<br />

R = radius of rotation<br />

The ratio of the centrifugal force to the weight is called the centrifugal number Nc: Nc = mR�2 /mg = R�2 /g (7-39)<br />

For liquid-to-liquid separation, the centrifugal number may be as high as 60,000 for<br />

certain tubular sedimentation designs. The mining industry is concerned with wear, so<br />

slurries are separated at centrifugal numbers smaller than 100.<br />

Cheremisinoff (1984) stated that the settling velocity of a particle in turbulent motion<br />

(Re > 500) in a centrifuge is Ks times as much as the free settling velocity, where


R<br />

Ks = 2�N � �� g<br />

The Reynolds number for the particle is calculated using the radial velocity:<br />

Re =<br />

For very fine particles with Re < 2, the migration is in laminar flow:<br />

For transition flow with 2 < Re < 500<br />

COMPONENTS OF SLURRY PLANTS<br />

FIGURE 7-44 Centrifugal separator. (Courtesy of Knelson Concentrators.)<br />

Ks = 4� 2N2 R<br />

� �<br />

4� 2N2R �<br />

g<br />

K s = � � 0.71<br />

(7-40)<br />

(7-41)<br />

(7-42)<br />

Consider a simple vertical centrifuge as in Figure 7-36. The solids in the <strong>slurry</strong> move<br />

toward the wall at a s<strong>pe</strong>ed u s toward the radius R w, while the liquid moves toward the axial<br />

feed tube at a s<strong>pe</strong>ed u L toward the radius R a. If the solids are at a volumetric concentration<br />

C V with a flow rate Q, the solids move at a s<strong>pe</strong>ed u s as<br />

Q s = 2�R 0Hu s = C vQ<br />

Separation will occur when u s > C vQ/2�R 0H.<br />

�2�RNd p<br />

��<br />

60�<br />

� g<br />

7.63


7.64 CHAPTER SEVEN<br />

Example 7-5<br />

A small centrifuge with a diameter of 150 mm is designed to handle 1.5 tons/hr of solids<br />

at a volumetric concentration of 40%. The density of the solids is 3000 kg/m3 . The height<br />

of the cone is 125 mm. Determine the minimum s<strong>pe</strong>ed of solids for separation from liquid.<br />

Solution<br />

Since the density is 3000 kg/m3 and the centrifuge handles 1500 kg/hr, the volume flow<br />

rate of solids is 0.5 m3 /hr, or 0.139 kg/s.<br />

For separation, us > 0.139/(2 × � × 0.15 × 0.125) and<br />

us > 1.18 m/s<br />

Considering the settling velocity of many particles, it is obvious that this centrifuge<br />

can handle the coarse particles found in certain mining <strong>systems</strong>.<br />

7-11 CONCLUSION<br />

To achieve many of the tasks described in this chapter, <strong>slurry</strong> must be transported from<br />

one point to another. This may be done by gravity flow, by o<strong>pe</strong>n channel flow, or by<br />

pumping. The pump is the workhorse of <strong>slurry</strong> transportation and will be analyzed in the<br />

next two chapters.<br />

A lot of different equipment is used in the processing of mineral ores. These were reviewed<br />

in this chapter more in terms of their place in the <strong>slurry</strong> circuit. The <strong>pe</strong>rformance<br />

of the equipment de<strong>pe</strong>nds on many factors such as pro<strong>pe</strong>r sizing and the characteristics of<br />

rocks and soils that too often cause extensive wear. The materials selected for processing<br />

by such equipment will be examined in Chapter 10, as they are also used as criteria in the<br />

manufacture of pumps.<br />

7-12 NOMENCLATURE<br />

A Area of flow across the pro<strong>pe</strong>ller<br />

c Blade chord<br />

C1, C2, C3 Coefficients of a hydrocyclone<br />

CD Drag coefficient<br />

CL Lift coefficient<br />

CQ Flow coefficient<br />

Cp Power coefficient<br />

Cr80 CVL d50 D<br />

d80 of the output wet ground rocks<br />

Volume fraction of liquid phase in a <strong>slurry</strong> tank<br />

d50 cut point of a hydrocyclone<br />

Drag force<br />

Di Conduit diameter (m)<br />

Din Diameter of mixer in inches<br />

Dimp Mixer im<strong>pe</strong>ller diameter<br />

Dm Mill diameter in meters<br />

Dmus Mill diameter in inches<br />

DT Mixer tank diameter<br />

e Natural number<br />

E1 Dry grinding factor


E2 Factor for o<strong>pe</strong>n circuit grinding to be expressed in terms of the final classification<br />

of solids<br />

E3 Mill diameter factor<br />

E4 Oversize feed factor for grinding<br />

E5 Fineness factor for ground or crushed particles<br />

E6 Reduction ratio factor for ball or rod mills<br />

E7 Low reduction ratio factor for ball or rod mills<br />

E8 Correction factor for rod mills<br />

E9 Correction factor for rubber-lined mills<br />

Fe80 Feop fw d80 of the feed rocks<br />

Optimum size of feed to a ball or rod mill<br />

Correlation factor for a mixer between design settling velocity and terminal<br />

velocity of solids<br />

g Acceleration due to gravity (9.8 m/s2 )<br />

H Height of mixer above bottom of tank<br />

HP Horsepower<br />

L Lift force<br />

n Number of im<strong>pe</strong>llers<br />

N Rotational s<strong>pe</strong>ed in rev/min<br />

P Power<br />

Q Flow rate (m3 /s)<br />

Rc Recovery of underflow from a cyclone<br />

Re Reynolds number<br />

ReB Reynolds number for a Bingham plastic, using the coefficient of rigidity for<br />

viscosity<br />

Rr material reduction ratio in a grinding circuit<br />

S Swirling number<br />

T Thrust force<br />

Uslip Slip s<strong>pe</strong>ed between liquid and solids in a mixer<br />

V Average velocity of flow (m/s)<br />

Vt Terminal velocity of solids<br />

W Consumed power for wet grinding<br />

Greek letters<br />

� Angle of incidence<br />

� Void fraction<br />

� Wet grinding factor<br />

� m<br />

Density of <strong>slurry</strong> mixture (kg/m 3 or dlugs/ft 3 )<br />

�s Density of solids in mixture (kg/m3 or dlugs/ft3 )<br />

� Factor of energy dissipation before the hydraulic jump in a free fall<br />

� Concentration by volume in decimal points<br />

� Shear strain<br />

� Pythagoras number (ratio of circumference of a circle to its diameter)<br />

� Duration of the shear for a time-de<strong>pe</strong>ndent fluid<br />

� Density<br />

� 0<br />

Yield stress for a Bingham plastic<br />

� Kinematic viscosity (usually expressed in Pascal-seconds or poise)<br />

� Angular velocity of particle<br />

Subscripts<br />

L Liquid<br />

m Mixture<br />

COMPONENTS OF SLURRY PLANTS<br />

7.65


7.66 CHAPTER SEVEN<br />

p Particle<br />

s Solids<br />

7-13 REFERENCES<br />

Arterburn, R. A. 1982. The sizing and selection of hydrocyclones. In Design and Installation of<br />

Communution Circuits, A. L. Mular and G. V. Jergensen (Eds.). New York: Society of Mining<br />

Engineers.<br />

Bond, F. C. 1952. Third theory of comminution. Trans. AIME, 193, 484.<br />

Burgess, K. E. and B. Abulnaga. 1991. The application of finite element analysis of Warman pumps<br />

and process equipment. Pa<strong>pe</strong>r presented at the Fifth International Conference on Finite Element<br />

Analysis, University of Sydney, Sydney, Australia.<br />

Cheremisinoff, N. P. 1984. Pocket Handbook for Solid–Liquid Separations. Houston: Gulf Publishing.<br />

Dickey, D. S. and J. G. Fenic. 1976. Dimensional analysis for fluid agitation <strong>systems</strong>. Chemical Engineering<br />

Elliott, A. J. 1991. Solids, communition, and grading. In Slurry Handling, edited by N. P. Brown and<br />

N. I. Heywood. New York: Elsevier Applied Sciences.<br />

Gates, L. E., J. R. Morton, and P. L. Fondy. 1976. Selecting agitator <strong>systems</strong> to sus<strong>pe</strong>nd solids in liquids.<br />

Chemical Engineering, May 24.<br />

Holmes, J. A. 1957. A contribution to the study of comminution, a modified form of Kick’s law.<br />

Trans. Inst. Chem. Engrs., 35, 125–156.<br />

Mular, A. L. and N. A. Jull. 1978. The selection of cyclone classifiers, pumps, and pump boxes for<br />

grinding circuits. In Mineral Processing Plant Design, A. L. Mular and R. B. Bhappu (Eds.).<br />

New York: Society of Mining Engineers.<br />

Oldshue, J. Y. 1983. Fluid Mixing Technology. New York: Chemical Engineering.<br />

Stephiewski, W. Z. and C. N. Keys. 1984. Rotary-Wing Aerodynamics. New York: Dover Publications.<br />

Stone, R. 1971. Ty<strong>pe</strong>s and costs of grinding equipment for solid waste water carriage. Pa<strong>pe</strong>r 19 in<br />

Advances in Solid–Liquid Flow in Pi<strong>pe</strong>s and Its Applications, edited by I. Zandi. New York:<br />

Pergamon Press, pp. 261–269:<br />

DENVER-SALA. 1995. Selection Guide for Process Equipment. Colorado Springs: Svedala Industries.<br />

Wasp, E. J., J. P. Kenny, and R. L. Gandhi. 1977. Solid–Liquid Flow Slurry Pi<strong>pe</strong>line Transportation.<br />

Aedermannsdorf, Switzerland: Trans Tech Publications.<br />

Weisman, J., and I. E. Efferding. 1960. Sus<strong>pe</strong>nsion of slurries by mechanical mixers. Am. Inst.<br />

Chem. Eng. Journal, 6, 419–426.<br />

Further readings<br />

Su, Y. S., and F. A. Holland. 1968. Agitation and mixing of non-Newtonian fluids. Chem. &<br />

Process. Eng., 49, 77–79.<br />

Turner, H. E., and H. E. McCarthy. 1965. Fundamental analysis of <strong>slurry</strong> grinding. AIChE, 15,<br />

581–584.


CHAPTER 4<br />

HETEROGENEOUS FLOWS<br />

OF SETTLING SLURRIES<br />

4-0 INTRODUCTION<br />

A practical engineer sifting through the literature on <strong>slurry</strong> flows would be astonished by<br />

the number of different equations. Since the work of the French scientists Durand and<br />

Condolios in 1952, and British scientists Newitt et al. in 1955, engineers and scientists<br />

have continued to develop new equations for deposition velocity and friction losses.<br />

This chapter reviews the evolution of models of Newtonian <strong>slurry</strong> flows from wellgraded<br />

and uniform particle sizes to complex mixtures of coarse and fine particles. For<br />

this purpose, some equations are listed in a historical context, to demonstrate their evolution<br />

to the reader. The sheer number of equations demonstrates how complicated heterogeneous<br />

flows are.<br />

A number of factors interact in a horizontal pi<strong>pe</strong>. The flow takes the form of different<br />

regimes, and includes everything from a simple stationary bed at low s<strong>pe</strong>ed to a pseudohomogeneous<br />

flow at high s<strong>pe</strong>eds. For each regime, equations have been develo<strong>pe</strong>d over<br />

the years to account for the mean particle diameter, the diameter of the conduit, the density<br />

of the particles, their drag coefficient, the s<strong>pe</strong>ed of flow, etc. As a result, there are many<br />

angles from which a heterogeneous flow of settling solids can be examined. The followers<br />

of Durand and Condolios put great emphasis on the drag coefficient of the solid particles,<br />

whereas the followers of Newitt prefer to focus on the terminal velocity. As we have<br />

clearly demonstrated in Chapter 3, the drag coefficient and the terminal velocity are interrelated.<br />

Examining a flow of <strong>slurry</strong> is often an exercise of playing with a murky liquid in<br />

which very little can be seen. Sand does not behave like coal and there is not a single universal<br />

law that may apply to the transportation of solids by liquids. It is therefore important<br />

to rely on database, historical information, and empirical data.<br />

In the past, engineers have tried to simplify the complexity of <strong>slurry</strong> flows by defining<br />

certain transition velocities. With the use of modern research tools, there is an emerging<br />

approach of rejecting the concept of an abrupt change from one state of flow to another,<br />

and a tendency to consider such a change over a band of the s<strong>pe</strong>ed. Different approaches<br />

have been develo<strong>pe</strong>d to examine the mixture of coarse and fine particles from su<strong>pe</strong>rimposed<br />

layers to two-layer models.<br />

This book is intended for engineers, and various examples are included in the text. The<br />

purpose of such examples is to simplify the use of complex equations. With modern <strong>pe</strong>rsonal<br />

computers, which use simple languages such as quick basic, an engineer can efficiently<br />

calculate friction losses for a heterogeneous <strong>slurry</strong> flow.<br />

4.1


4.2 CHAPTER FOUR<br />

4-1 REGIMES OF FLOW OF A<br />

HETEROGENEOUS MIXTURE IN<br />

HORIZONTAL PIPE<br />

The history of <strong>slurry</strong> pi<strong>pe</strong>lines was briefly presented in Section 1-10 of Chapter 1. Two<br />

schools are credited for laying the foundations of modern hydro-transport engineering;<br />

SOGREAH in France, and the British Hydro-mechanic Research Association of the United<br />

Kingdom. Starting in 1952, Durand and Condolios of SOGREAH published a number<br />

of studies on the flow of sand and gravel in pi<strong>pe</strong>s up to 900 mm (35.5 in) in diameter.<br />

Based on the s<strong>pe</strong>cific gravity of particles with a magnitude of 2.65, they proposed to divide<br />

the flows of nonsettling slurries in horizontal pi<strong>pe</strong>s into four categories based on average<br />

particle size as follows:<br />

� Homogeneous sus<strong>pe</strong>nsions for particles smaller than 40 �m (mesh 325)<br />

� Sus<strong>pe</strong>nsions maintained by turbulence for particle sizes from 40 �m (mesh 325) to<br />

0.15 mm (mesh 100)<br />

� Sus<strong>pe</strong>nsion with saltation for particle sizes between 0.15 mm (mesh 100) and 1.5 mm<br />

(mesh 11)<br />

� Saltation for particles greater than 1.5 mm (mesh 11)<br />

This initial classification was refined over the next 18 years by Newitt et al. (1955),<br />

Ellis and Round (1963), Thomas (1964), Shen (1970), and Wicks (1971). Due to the interrelation<br />

between particle sizes and terminal and deposition velocities, the original classification<br />

proposed by Durand has been modified to four flow regimes based on the actual<br />

flow of particles and their size. Referring to Figures 4-1 and 4-2, there are four main<br />

regimes of flow in a horizontal pi<strong>pe</strong><br />

1. Flow with a stationary bed<br />

2. Flow with a moving bed and saltation (with or without sus<strong>pe</strong>nsion)<br />

3. Heterogeneous mixture with all solids in sus<strong>pe</strong>nsion<br />

4. Pseudohomogeneous or homogeneous mixtures with all solids in sus<strong>pe</strong>nsion<br />

velocity<br />

Fully sus<strong>pe</strong>nded<br />

lenticular<br />

deposits<br />

stationary<br />

deposits<br />

with ripples<br />

sus<strong>pe</strong>nded with moving bed<br />

sus<strong>pe</strong>nded with saltation<br />

blocked<br />

pi<strong>pe</strong><br />

FIGURE 4-1 Flow regimes of heterogeneous flows in terms of s<strong>pe</strong>ed versus volumetric concentration<br />

(after Newitt et al., 1955).


particle size<br />

Flow with a stationary bed<br />

Two s<strong>pe</strong>cial cases shown in Figure 4-1 are not considered to be at the limits of these<br />

regimes of flows. They are lenticular deposits at very low s<strong>pe</strong>eds but low solid concentration,<br />

and blocked pi<strong>pe</strong>s at high solid concentration.<br />

Slurry flows that have some form of segregation or separation of solids in layers are<br />

called “heterogeneous” flows, whereas the slurries themselves are called “settling” slurries.<br />

4-1-1 Flow With a Stationary Bed<br />

When the <strong>slurry</strong> flow s<strong>pe</strong>ed is low, the bed thickens. As the fluid above the bed tries to<br />

move the solids by entrainment, they tend to roll and tumble. The particles with the lowest<br />

settling s<strong>pe</strong>ed move as an asymmetric sus<strong>pe</strong>nsion, whereas the coarser particles build<br />

up the bed. As the s<strong>pe</strong>ed drops even further, the pressure to maintain the flow becomes<br />

quite high and eventually the pi<strong>pe</strong> blocks up.<br />

Flow with saltation and asymmetric sus<strong>pe</strong>nsion occurs above the s<strong>pe</strong>ed of blockage.<br />

This means that the coarser particles “sand up,” whereas the finer particles continue to<br />

move. Certain tailing lines have exhibited this phenomenon. In fact, when a process plant<br />

is built with a tailing line too large to handle the initial flow, the o<strong>pe</strong>rator may choose to let<br />

the bottom of the pi<strong>pe</strong> sand up to reduce the effective cross-sectional area of the pi<strong>pe</strong>. This<br />

principle has been successfully applied to pi<strong>pe</strong>lines in a variety of countries. Saltation can<br />

eventually lead to blockage of a pi<strong>pe</strong>. It may result in a number of problems, such as water<br />

hammer, wear, and freezing in cold climates. Most engineering s<strong>pe</strong>cifications require that<br />

the pi<strong>pe</strong>line be designed to o<strong>pe</strong>rate at s<strong>pe</strong>eds higher than those associated with saltation.<br />

4-1-2 Flow With a Moving Bed<br />

HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

Flow with a moving bed<br />

with or without sus<strong>pe</strong>nsions<br />

Heteregeneous flow with all<br />

solids in sus<strong>pe</strong>nsion<br />

Flow as a homogeneous or<br />

pseudohomogeneous sus<strong>pe</strong>nsion<br />

Mean velocity<br />

FIGURE 4-2 Flow regimes of heterogeneous flows in terms of particle size versus mean velocity<br />

(after Shen, 1970).<br />

When the s<strong>pe</strong>ed of the flow is low and there are a large number of coarse particles, the<br />

bed moves like desert sand dunes. The top particles are entrained in the moving fluid<br />

4.3


4.4 CHAPTER FOUR<br />

above the bed. Consequently, the up<strong>pe</strong>r layers of the bed move faster than the lower layers<br />

in a horizontal pi<strong>pe</strong>. If the mixture were composed of a wide range of particles with<br />

different sizes and settling velocities, the bed would be composed of the particles with the<br />

highest settling s<strong>pe</strong>ed. Particles with a moderate settling s<strong>pe</strong>ed are maintained in an asymmetric<br />

sus<strong>pe</strong>nsion, with most particles concentrated in the lower half of the pi<strong>pe</strong>, whereas<br />

the particles with the lowest settling s<strong>pe</strong>ed move as a symmetric sus<strong>pe</strong>nsion<br />

4-1-3 Sus<strong>pe</strong>nsion Maintained by Turbulence<br />

As the flow s<strong>pe</strong>ed increases, turbulence is sufficient to lift more solids. All particles move<br />

in an asymmetric pattern with the coarsest at the bottom of a horizontal pi<strong>pe</strong> covered with<br />

su<strong>pe</strong>rimposed layers of medium- and fine-sized particles. Many particles may strike the<br />

bottom of the pi<strong>pe</strong> and rebound. Wear on the bottom of the pi<strong>pe</strong> must be taken into account<br />

in maintenance schedules and the pi<strong>pe</strong>s must be rotated at intervals suggested by<br />

the <strong>slurry</strong> engineer in order to maintain an even wear pattern of the internal wall of the<br />

pi<strong>pe</strong>. Although the flow is not symmetric, from the point of view of power consumption,<br />

this regime may be the most economical for transporting a certain mass of solids.<br />

Wilson (1991) calls all flows below V 3 fully stratified flows and all flows above V 3<br />

fully sus<strong>pe</strong>nded flows. The transition from fully stratified to fully sus<strong>pe</strong>nded flows is considered<br />

by this author to be fairly complex and should be represented by sigmoid or ogee<br />

curves. It is a transition over a range of the s<strong>pe</strong>ed and not an abrupt transition at a single<br />

value of the s<strong>pe</strong>ed. The work of Wilson and colleagues will be examined in Section 4-4-5.<br />

4-1-4 Symmetric Flow at High S<strong>pe</strong>ed<br />

At s<strong>pe</strong>eds in excess of 3.3 m/s (10 ft/s), all solids may move in a symmetric pattern (but<br />

not necessarily uniformly). Sometimes this flow is called pseudohomogeneous because of<br />

its symmetry around the pi<strong>pe</strong> axis. Power consumption is a linear relationship of the static<br />

head multiplied by the velocity, but is proportional to the cube of velocity needed to<br />

overcome friction losses. Power consumption in pseudohomogeneous mixtures of coarse<br />

and fine particles may be excessive for long pi<strong>pe</strong>lines. Pseudohomogeneous mixtures of<br />

fine or ultrafine particles may occur at s<strong>pe</strong>eds as low as 1.52 m/s (5 ft/s). One definition of<br />

fine and coarse particles was explained Govier and Aziz (1972), who proposed the following:<br />

� Ultrafine particles: d p < 10 �m (mesh 1250), where gravitational forces are negligible.<br />

� Fine particles: 10 �m < d p < 100 �m (mesh 1250 < d p < mesh 140), usually carried fully<br />

sus<strong>pe</strong>nded but subject to concentration gradients and gravitational forces.<br />

� Medium sized particles: 100 �m < d p < 1000 �m (mesh 240 < d p < mesh 15), will<br />

move with a deposit at the bottom of the pi<strong>pe</strong> and with a concentration gradient.<br />

� Coarse particles: 1000 �m < d p < 10,000 �m (0.039 in < d p < 0.394 in). These are seldom<br />

fully sus<strong>pe</strong>nded and form deposits on the bottom of the pi<strong>pe</strong>.<br />

� Ultracoarse particles are larger than 10 mm (0.4 in). These particles are transported as<br />

a moving bed on the bottom of the pi<strong>pe</strong>.<br />

Considering particle sizes while ignoring their density is meaningless. Practical engineers<br />

do shift the boundaries between different sizes based on the density of the particles.<br />

There is no question that beads of high-density polyethylene will behave differently than


sand particles with the same average diameter because the former is lighter than water<br />

while the latter is 2.65 times heavier than water.<br />

4-2 HOLD-UP<br />

HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

The previous section describes how different layers of solids move with different s<strong>pe</strong>eds,<br />

from the bottom, coarser particles, to the finer particles at the top of the horizontal pi<strong>pe</strong>.<br />

The theory of hold-ups complicates this process, however. Hold-ups are due to velocity<br />

slip of layers of particles of larger sizes, particularly in the moving bed flow pattern.<br />

Newitt et al. (1962) conducted s<strong>pe</strong>ed measurements of a <strong>slurry</strong> mixture in a horizontal<br />

pi<strong>pe</strong>. In the case of light Plexiglas pi<strong>pe</strong>, zircon or fine sand did not result in local slip; particles<br />

and water moved at the same s<strong>pe</strong>ed. However, for coarse sand and gravel, they observed<br />

asymmetric sus<strong>pe</strong>nsion and a sliding bed. They also observed that in the up<strong>pe</strong>r layers<br />

of the horizontal pi<strong>pe</strong>, the concentrations of larger particles were the same as for finer<br />

solids, but were marked by differences in the magnitude of the discharge rate of the lower<br />

layers.<br />

4.3 TRANSITIONAL VELOCITIES<br />

The four regimes of flow described in Section 4-1 can be represented by a plot of the<br />

pressure gradient versus the average s<strong>pe</strong>ed of the mixture (Figure 4-5). The transitional<br />

velocities are defined as<br />

� V 1: velocity at or above which the bed in the lower half of the pi<strong>pe</strong> is stationary. In the<br />

up<strong>pe</strong>r half of the pi<strong>pe</strong>, some solids may move by saltation or sus<strong>pe</strong>nsion.<br />

Ratio distanc e from bottom of pi<strong>pe</strong><br />

to the inner diameter (y/D ) I<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0<br />

0.03<br />

C<br />

v<br />

0.07<br />

0.10<br />

0.14<br />

0.05 0.1 0.15 0.20<br />

Discharge solids concentration Cy<br />

FIGURE 4-3 Distribution of concentration of solids in a pi<strong>pe</strong> versus average volumetric<br />

concentration.<br />

4.5


4.6 CHAPTER FOUR<br />

� V 2: velocity at or above which the mixture flows as an asymmetric mixture with the<br />

coarser particles forming a moving bed.<br />

� V 3 or V D: velocity at or above which all particles move as an asymmetric sus<strong>pe</strong>nsion<br />

and below which the solids start to settle and form a moving bed.<br />

� V 4: velocity at or above which all solids move as a symmetric sus<strong>pe</strong>nsion.<br />

Volumetric Concentration (%)<br />

30<br />

20<br />

10<br />

0<br />

0<br />

1 2 3 4 5<br />

Velocity (m/s)<br />

FIGURE 4-4 Simplified concept of particle distribution in a pi<strong>pe</strong> as a function of volumetric<br />

concentration and s<strong>pe</strong>ed.<br />

Pressure drop <strong>pe</strong>r unit of length<br />

Velocity (ft/sec)<br />

0 5 10 15 20<br />

1<br />

2<br />

<strong>slurry</strong><br />

3<br />

V V V V 1 2 3<br />

4<br />

stationary bed<br />

moving bed<br />

asymmetric flow<br />

direction of flow<br />

4<br />

water<br />

symmetric flow<br />

6<br />

S<strong>pe</strong>ed of flow<br />

FIGURE 4-5 Velocity regimes for heterogeneous <strong>slurry</strong> flows.


V 3 is effectively the deposition velocity, often called in the past the Durand velocity<br />

for uniformly sized coarse particles. It is no longer recommended that it be called the Durand<br />

velocity, as tests in the last 20 years have led to new equations that include the effects<br />

of particle size and composition of the <strong>slurry</strong>. The magnitude of the velocity de<strong>pe</strong>nds<br />

on the volumetric concentration (Figure 4-7).<br />

4-3-1 Transitional Velocities V 1 and V 2<br />

The transitional velocity V 1 is obviously not used for the o<strong>pe</strong>ration of <strong>slurry</strong> lines. It may<br />

be of interest in lab research, instrumentation, and monitor of start-up.<br />

The transitional velocity V 2 is determined individually from pressure measurements of<br />

the pressure gradient. The main focus of the tests is to determine the height of the bed and<br />

to derive a stratification ratio.<br />

Wilson (1970) develo<strong>pe</strong>d a model for the incipient motion of granular solids at V 2. He<br />

assumed a hydrostatic pressure exerted by the solids on the wall and proposed the following<br />

equation:<br />

1<br />

� �L<br />

� �� + �sin � – �P � – sin � cos � Rw ��s �� ��<br />

� L<br />

4<br />

� tan �r<br />

�s(S – 1) Cvb(sin � – cos �)g<br />

= ���<br />

(4-1)<br />

2<br />

where<br />

(�P/L) 2 = pressure gradient at 2<br />

� = half the angle subtended at the pi<strong>pe</strong> center due to the up<strong>pe</strong>r surface of the bed,<br />

in radians<br />

�s = coefficient of static friction of the solid particles against the wall of the pi<strong>pe</strong><br />

Rw = cross-sectional area of the bed divided by the bed width<br />

�r = angle of repose of the solid particles<br />

Cumulative passing<br />

(%)<br />

HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

100<br />

10<br />

FIGURE 4-6 Concept of d 50 by cumulative passing <strong>pe</strong>rcentage versus particle size.<br />

� Di<br />

0<br />

0 10 100 1000<br />

d 50<br />

Particle mesh diameter ( m)<br />

4.7


4.8 CHAPTER FOUR<br />

S = ratio of density of solids to density of liquid<br />

Cvb = volume fraction solids in the bed<br />

(When USCS units are used, express density in slugs/ft3 rather than lbm/ft3 ).<br />

For 0.7 mm (mesh 24) sand with water in a 90 mm (3.5 in) pi<strong>pe</strong>, Wilson measured � =<br />

0.35 and concluded that the assumptions of hydrostatic distribution of the granular pressure<br />

were correct.<br />

4-3-2 The Transitional Velocity V 3 or S<strong>pe</strong>ed for Minimum<br />

Pressure Gradient<br />

The transitional velocity V3 is extremely important because it is the s<strong>pe</strong>ed at which the<br />

pressure gradient is at a minimum. Although there is evidence that solids start to settle at<br />

slower s<strong>pe</strong>eds in complex mixtures, o<strong>pe</strong>rators and engineers often referred to transitional<br />

velocity as the s<strong>pe</strong>ed of deposition.<br />

Durand and Condolios (1952) derived the following equation for uniformly sized sand<br />

and gravel:<br />

VD = V3 = FL{2 · g · Di[(�s – �L)/�L]} 1/2 (4-2)<br />

where<br />

FL = is the Durand factor based on grain size and volume concentration<br />

V3 = the critical transition velocity between flow with a stationary bed and a heterogeneous<br />

flow<br />

Di = pi<strong>pe</strong> inner diameter (in m)<br />

g = acceleration due to gravity (9.81 m/s)<br />

�s = density of solids in a mixture (kg/m3 )<br />

�L = density of liquid carrier<br />

The Durand factor FL is typically represented in a graph for single or narrow graded<br />

particles, as in Figure 4-7 after the work of Durand (1953). However, since most slurries<br />

Durand Velocity Factor<br />

F<br />

D<br />

2.0<br />

1.0<br />

0<br />

C = 5%<br />

V<br />

C = 2%<br />

V<br />

C<br />

V<br />

= 15%<br />

C<br />

V<br />

= 10%<br />

0 1.0 2.0 3.0<br />

Particle diameter (mm)<br />

C<br />

V<br />

= 15%<br />

C<br />

V<br />

= 5%<br />

For single or narrow<br />

graded slurries<br />

Based on Schiller<br />

equation using d<br />

50<br />

FIGURE 4-7 Durand velocity factor versus particle size—comparison between the conventional<br />

values for single graded slurries and Schiller’s equation using d 50 for wide graded <strong>slurry</strong>.


are mixtures of particles of different sizes, this plot is considered too conservative. The<br />

Durand velocity factor has been refined by a number of authors.<br />

In an effort to represent more diluted concentrations, Wasp et al. (1970) proposed including<br />

a ratio between the particle diameter and the pi<strong>pe</strong>line diameter. Wasp proposed<br />

the use of a modified factor F� L so that<br />

1/2<br />

VD = V3 = F� L�2gDi � � � 1/6<br />

�S – �L dp (4-3)<br />

Schiller and Herbich (1991) proposed the following equation for the Durand velocity<br />

factor based on the d 50 of the particles:<br />

F L = {(1.3 × C v 0.125 )[1 – exp (–6.9 d50)]} (4-4)<br />

where d 50 is expressed in mm.<br />

Some reference books define a Froude number as Fr = F L · �2�. The particle size d 50<br />

is the statistically determined particle size below which half (or 50%) would be equal or<br />

smaller to that set size. The following example illustrates the concept of d 50.<br />

Example 4-1<br />

A sample of <strong>slurry</strong> is sieved for particle size. The data is collected in the laboratory (see<br />

Table 4-1). Plot the data on a logarithmic graph and determine the d50. Solution<br />

The data is plotted in Figure 4-6; the d50 is determined to be 145 �m.<br />

Example 4-2<br />

A <strong>slurry</strong> mixture has a d 50 of 300 �m. The <strong>slurry</strong> is pum<strong>pe</strong>d in a 30 in pi<strong>pe</strong> with an<br />

ID of 28.28�. The volumetric concentration is 0.27. Using Equations 4-4 and 4-2, determine<br />

the s<strong>pe</strong>ed of deposition for a sand–water mixture if the s<strong>pe</strong>cific gravity of sand is<br />

2.65.<br />

Solution in SI Units<br />

From Equation 4-4:<br />

F L = (1.3 × 0.27 0.125 )(1-exp (–6.9 × 0.3))<br />

F L = 1.1 × 0.8738<br />

F L = 0.964<br />

From Equation 4-2 the deposition velocity is<br />

TABLE 4-1 Data for Example 4-1<br />

HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

� �L<br />

V 3 = 0.964 (2 × 9.81 × 28.25 × 0.0254 × 1.65) 1/2<br />

V 3 = 4.64 m/s<br />

Particle size (�m) 425 300 212 150 106 75 53 45 38 –38<br />

Cumulative 97.2 87.1 68.3 51.3 35.9 20.5 14.5 11.8 10.8 —<br />

passing (%)<br />

� Di<br />

4.9


4.10 CHAPTER FOUR<br />

Solution in USCS Units<br />

From Equation 4-4:<br />

FL = (1.3 × 0.270.125 )(1 – exp (–6.9 × 0.3))<br />

FL = 1.1 × 0.8738<br />

FL = 0.964<br />

From Equation 4-2 the deposition velocity is<br />

V3 = 0.964 (2 × 32.2 × 28.25/12 × 1.65) 1/2<br />

V 3 = 15.25 ft/sec<br />

Various curves have been published for the magnitude of F L. They are often plotted<br />

for a single graded size and use difficult to read logarithmic scales. For the sake of accuracy,<br />

Table 4-2 tabulates the magnitude of F L between 0.08 mm < d 50 < 5mm on the basis<br />

TABLE 4-2 The Coefficient F L Based on Schiller’s Equation Using the d 50 of the<br />

Particles for Particles Between 0.080 and 5 mm for Volumetric Concentration up to<br />

30%. F L = {(1.3 × C v 0.125 )[1 – exp(–6.9 d50)]}<br />

d 50 (mm) C V = 0.05 C V = 0.10 C V = 0.15 C V = 0.20 C V = 0.25 C V = 0.30<br />

0.08 0.379 0.414 0.435 0.451 0.464 0.474<br />

0.10 0.446 0.486 0.511 0.530 0.545 0.557<br />

0.12 0.503 0.549 0.577 0.599 0.616 0.630<br />

0.14 0.554 0.604 0.635 0.658 0.677 0.693<br />

0.16 0.598 0.652 0.686 0.711 0.731 0.748<br />

0.18 0.636 0.693 0.729 0.756 0.777 0.795<br />

0.20 0.669 0.730 0.768 0.796 0.818 0.837<br />

0.25 0.735 0.801 0.843 0.874 0.898 0.919<br />

0.30 0.781 0.852 0.896 0.929 0.955 0.977<br />

0.35 0.814 0.888 0.934 0.968 0.995 1.018<br />

0.40 0.837 0.913 0.961 0.996 1.024 1.048<br />

0.45 0.854 0.931 0.980 1.015 1.044 1.068<br />

0.50 0.866 0.944 0.993 1.029 1.058 1.083<br />

0.55 0.874 0.953 1.002 1.039 1.069 1.093<br />

0.60 0.880 0.959 1.009 1.046 1.076 1.101<br />

0.65 0.884 0.964 1.014 1.051 1.081 1.106<br />

0.70 0.887 0.967 1.017 1.055 1.084 1.109<br />

0.75 0.889 0.969 1.020 1.057 1.087 1.112<br />

0.80 0.890 0.971 1.021 1.059 1.089 1.114<br />

0.85 0.891 0.972 1.023 1.060 1.090 1.115<br />

0.90 0.892 0.973 1.023 1.061 1.091 1.116<br />

1.00 0.893 0.974 1.0245 1.062 1.092 1.1172<br />

1.5 0.8939 0.9748 1.0255 1.063 1.0931 1.1183<br />

2 0.8940 0.9749 1.0255 1.063 1.0932 1.1184<br />

2.5 0.8940 0.9749 1.0255 1.063 1.0932 1.1184<br />

3.0 0.8940 0.9749 1.0255 1.063 1.0932 1.1184<br />

3.5 0.8940 0.9749 1.0255 1.063 1.0932 1.1184<br />

4.0 0.8940 0.9749 1.0255 1.063 1.0932 1.1184<br />

5.0 0.8940 0.9749 1.0255 1.063 1.0932 1.1184


HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

of Schiller’s equation. The magnitude of FL based on d50 is smaller than values published<br />

in the literature for single graded <strong>slurry</strong> mixtures (lab mixtures using a uniform size of<br />

particles). A number of authors have confirmed that this is the case (Warman International<br />

Inc., 1990).<br />

In order to compare the conventional magnitude of FL based on single and narrow<br />

graded particles to the Schiller equation, both ranges of FL are plotted in Figure 4-7.<br />

With a more complex approach that takes into account the actual viscosity of the <strong>slurry</strong><br />

mixture and the density of the particles, Gillies et al. (1999) develo<strong>pe</strong>d an equation for<br />

the Froude number F in terms of the Archimedean number (which we will discuss in Section<br />

4-4-5 for stratified coarse flows):<br />

4<br />

3 Ar = d p �L(�s – �L)g (4-5)<br />

To estimate the deposition velocity V3, Gilles et al. (1999) develo<strong>pe</strong>d an equation for<br />

the Froude number based on the Archimedean number:<br />

Fr = aArb (4-6)<br />

where<br />

Fr = FL · �2�<br />

For Ar > 540, a = 1.78, b = –0.019<br />

For 160 < Ar < 540, a = 1.19, b = 0.045<br />

For 80 < Ar < 60, a = 0.197, b = 0.4<br />

For Ar < 80, the Wilson and Judge (1976) equation can be used, which expressed the<br />

Froude number as<br />

Fr = (�2�)�2.0 + 0.30 log dp 10� �� (4-7)<br />

This correlation is useful in the range of 10 –5 < (dp/DiCD) < 10 –3 �<br />

DiCD .<br />

To determine the drag coefficient, the actual density of the liquid should be used,<br />

whereas the viscosity should be corrected for the presence of fines.<br />

Example 4-3<br />

Water at a viscosity of 0.0015 Pa · s (0.0000313 slugs/ft-sec) is used to transport sand<br />

with an average particle diameter of 300 �m (0.0118 inch). The volumetric concentration<br />

is 0.27. The pi<strong>pe</strong>’s inner diameter is 717 mm (28.35�). Using the Gilles equation (Equation<br />

4-6), determine the deposition velocity if the s<strong>pe</strong>cific gravity of sand is 2.65. Assume<br />

CD = 0.45.<br />

Solution in SI Units<br />

d 50<br />

� Di<br />

� 3�L 2<br />

= = 0.4 × 10 –3<br />

0.3<br />

�<br />

717<br />

Iteration 1<br />

Assuming C D > 10 –3 , by the Wilson and Judge correlation (Equation 4-7):<br />

Fr = (�2�)�2.0 + 0.30 log 0.003<br />

10����� 0.717 × 0.45<br />

Fr = 1.54<br />

FL = Fr/�2� = 1.54/�2� = 1.09<br />

4.11


4.12 CHAPTER FOUR<br />

The s<strong>pe</strong>cific gravity of the mixture is determined as:<br />

Sm = Cv(Ss – SL) + SL = 0.27 (2.65 – 1) + 1 = 1.446<br />

VD = FL�[2�g�D� i(��� s/��� L�–� 1�]� = 4.82 m/s<br />

Iteration 2<br />

Ar = = 258.98<br />

for 160 < Ar < 540, a = 1.19, b = 0.045.<br />

From equation 4.6:<br />

Fr = aArb = 1.19 · 258.980.045 = 1.53<br />

FL = F/�2� = 1.53/�2� = 1.082<br />

VD = FL[2gDi(�s/�L – 1)] 0.5<br />

4 × 9.81 (3 × 10 –4 ) 3 × 1000 (1650)<br />

����<br />

3(1.5 × 10 –3 ) 2<br />

Solution in USCS Units<br />

V D = 1.082[2 · 9.81 · 0.717 · (1.65)] 0.5 = 5.21m/s<br />

d 50<br />

� Di<br />

= = 0.4 × 10 –3<br />

0.00118<br />

�<br />

28.23<br />

Iteration 1<br />

Assuming C D > 10 –3 , by the Wilson and Judge correlation (Equation 4-7):<br />

Fr = (�2�)�2.0 + 0.30 log 0.00118<br />

10����� Fr = 1.54<br />

FL = 1.54/2 = 1.09<br />

The s<strong>pe</strong>cific gravity of the mixture is determined as:<br />

Sm = Cv(Ss – SL) +SL = 0.27(2.65 – 1) + 1 = 1.446<br />

VD = 1.09[2 · 32.2 · (28.23/12) (2.65 – 1)] 0.5<br />

VD = 17.23 ft/sec<br />

Iteration 2<br />

The particles’ diameter is 0.984 · 10 –3 ft<br />

The density of water is 1.93 slugs/ft3 The density of sand is 5.114 slugs/ft3 Water dynamic viscosity is 0.0000313 slugs/ft-sec<br />

4(0.984 · 10<br />

Ar = = 259<br />

–3 ) 3 28.23 × 0.45<br />

× 1.93(5.114 – 1.93) · 32.2<br />

������<br />

3(0.0000313) 2<br />

for 160 < Ar < 540, a = 1.19, b = 0.045.<br />

From equation 4.6,<br />

Fr = aArb = 1.19 · 258.980.045 = 1.53<br />

FL = 1.53/�2� = 1.082<br />

VD = FL[2gDi(�s/�L – 1)] 0.5<br />

V D = 1.082[2 · 32.2 · 2.35 · (1.65)] 0.5 = 17.1 ft/s


HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

4.13<br />

The solution by the Gilles equation is within the limits set by Schiller in Example 4-2.<br />

In these two different examples, we applied two different formulae but obtained consistent<br />

results. This demonstrates the sensitivity of approaches to equations derived from<br />

empirical equations. It may be necessary sometimes try to solve a problem using two different<br />

equations, and to use common sense when similar results are obtained.<br />

Table 4-3 presents values of the Archimedean number, the resultant magnitude of the<br />

factor F L for particles d 50 in the range of 0.08 to 50 mm for a s<strong>pe</strong>cific gravity of 1.5,<br />

which is typical of coal-based mixtures. Most coals may be pum<strong>pe</strong>d with different sizes<br />

of particles as discussed in Chapter 11. The viscosity may be due to the presence of cer-<br />

TABLE 4-3 The Coefficient F L Based on Gilles Equation for Particles Between 0.080<br />

and 50 mm of S<strong>pe</strong>cific Gravity of 1.500 (e.g., Coal) as a Function of Viscosity<br />

� = 1 cP � = 5 cP � = 10 cP<br />

_____________________ _______________________ _______________________<br />

Archimedean Archimedean Archimedean<br />

d 50 (mm) number Ar F L number Ar F L number Ar F L<br />

0.08 3.35 Eqn 4-7 0.13 Eqn 4-7 0.033 Eqn 4-7<br />

0.10 6.54 Eqn 4-7 0.26 Eqn 4-7 0.065 Eqn 4-7<br />

0.12 11.3 Eqn 4-7 0.45 Eqn 4-7 0.113 Eqn 4-7<br />

0.14 17.9 Eqn 4-7 0.72 Eqn 4-7 0.18 Eqn 4-7<br />

0.16 26.8 Eqn 4-7 1.07 Eqn 4-7 0.27 Eqn 4-7<br />

0.18 38.1 Eqn 4-7 1.53 Eqn 4-7 0.38 Eqn 4-7<br />

0.20 52.3 Eqn 4-7 2.1 Eqn 4-7 0.52 Eqn 4-7<br />

0.25 102 0.89 4.1 Eqn 4-7 1.02 Eqn 4-7<br />

0.30 177 1.062 7.1 Eqn 4-7 1.77 Eqn 4-7<br />

0.35 280 1.084 11.2 Eqn 4-7 2.80 Eqn 4-7<br />

0.40 419 1.104 16.75 Eqn 4-7 4.19 Eqn 4-7<br />

0.45 596 1.420 23.8 Eqn 4-7 5.96 Eqn 4-7<br />

0.50 818 1.43 32.7 Eqn 4-7 8.18 Eqn 4-7<br />

0.55 1088 1.437 43.5 Eqn 4-7 10.9 Eqn 4-7<br />

0.60 1413 1.445 56.51 Eqn 4-7 14.1 Eqn 4-7<br />

0.65 1796 1.451 72 Eqn 4-7 18 Eqn 4-7<br />

0.70 2243 2.457 89.7 0.84 22.4 Eqn 4-7<br />

0.75 2579 1.463 110.4 0.914 27.6 Eqn 4-7<br />

0.80 3348 1.469 134 0.99 33.5 Eqn 4-7<br />

0.85 4016 1.474 161 1.058 40 Eqn 4-7<br />

0.90 4768 1.478 191 1.066 48 Eqn 4-7<br />

1.00 6540 1.487 262 1.081 65 Eqn 4-17<br />

2.00 52320 1.547 2093 1.455 523 1.12<br />

3.00 176580 1.583 7063 1.489 1765 1.45<br />

4.00 418560 1.610 16742 1.514 4185 1.475<br />

5.00 817500 1.63 32700 1.533 8175 1.494<br />

6.00 1415640 1.647 56505 1.55 14126 1.51<br />

8.00 3348480 1.674 133939 1.575 33485 1.534<br />

10.00 6540000 1.696 261600 1.595 65400 1.554<br />

20.00 5.23 × 10 7 1.764 2092800 1.66 523200 1.616<br />

30.00 17.7 × 10 8 1.805 7063202 1.698 1765800 1.654<br />

40.00 41.86 × 10 8 1.835 16742404 1.726 4185601 1.682<br />

50.00 81.75 × 10 8 1.859 32700008 1.749 81750020 1.703


4.14 CHAPTER FOUR<br />

tain fines, as with <strong>pe</strong>at coals or degradation of the coal during pumping over long distances,<br />

or the use of a heavy medium such as magnetite at high concentration as a carrier<br />

for coal in a water mixture.<br />

Table 4-4 presents values of the factor F L for particles d 50 in the range of 0.08 to 50<br />

mm for a s<strong>pe</strong>cific gravity of 2.65, which is typical of sand and tar-sand-based mixtures.<br />

The largest particles are often found in tar sand applications, with some contribution of<br />

the tar or oil to viscosity. In this table, there was no need to present the Archimedean<br />

number, as this was demonstrated in the previous table.<br />

Newitt et al. (1955) preferred to express the s<strong>pe</strong>ed of transition between “saltation”<br />

flow and heterogeneous flow in terms of the terminal velocity of particles (previously discussed<br />

in Chapter 3):<br />

V 3 = 17 V t<br />

(4.8)<br />

The reader should refer to Equation 3-18, which corrects the terminal velocity of a single<br />

particle to a mass of particles at higher volumetric concentration. Although Equation<br />

4-8 has served as the basis of many models, we will later discuss the recent corrections<br />

proposed by Wilson et al. (1992).<br />

The approach to obtain the magnitude of V 3 is basically to conduct a test and measure<br />

pressure drop <strong>pe</strong>r unit length of pi<strong>pe</strong>. V 3 is considered to occur at the minima, or the point<br />

of minimum pressure drop. W. E. Wilson (1942) expressed the pressure gradient of noncolloidal<br />

solids by referring to clean water and by proposing a correction to the<br />

Darcy–Weisbach equation (discussed in Chapter 2). He expressed the consumed power<br />

due to friction by the following equation:<br />

FIGURE 4-8 These taconite tailings must be pum<strong>pe</strong>d above a deposit velocity of 13 ft/s in<br />

14� pi<strong>pe</strong> due to the size of the particles.


HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

TABLE 4-4 The Coefficient F L Based on Gilles Equation for Particles Between 0.080<br />

and 50 mm of S<strong>pe</strong>cific Gravity of 2.65 (e.g., Sands and Oil Sands) as a Function of<br />

Viscosity<br />

d50 (mm) � = 1 cP, FL � = 5 cP, FL � = 10 cP, FL 0.08 Eqn 4-7 Eqn 4-7 Eqn 4-7<br />

0.10 Eqn 4-7 Eqn 4-7 Eqn 4-7<br />

0.12 Eqn 4-7 Eqn 4-7 Eqn 4-7<br />

0.14 Eqn 4-7 Eqn 4-7 Eqn 4-7<br />

0.16 0.837 Eqn 4-7 Eqn 4-7<br />

0.18 0.964 Eqn 4-7 Eqn 4-7<br />

0.20 1.061 Eqn 4-7 Eqn 4-7<br />

0.25 1.093 Eqn 4-7 Eqn 4-7<br />

0.30 1.421 Eqn 4-7 Eqn 4-7<br />

0.35 1.433 Eqn 4-7 Eqn 4-7<br />

0.40 1.444 Eqn 4-7 Eqn 4-7<br />

0.45 1.454 0.8 Eqn 4-7<br />

0.50 1.462 0.906 Eqn 4-7<br />

0.55 1.470 1.016 Eqn 4-7<br />

0.60 1.478 1.065 Eqn 4-7<br />

0.65 1.485 1.076 Eqn 4-7<br />

0.70 1.491 1.087 Eqn 4-17<br />

0.75 1.497 1.097 0.847<br />

0.80 1.502 1.107 0.915<br />

0.85 1.507 1.116 0.984<br />

0.90 1.512 1.423 1.054<br />

1.00 1.521 1.431 1.072<br />

2.00 1.583 1.489 1.450<br />

3.00 1.620 1.524 1.484<br />

4.00 1.647 1.549 1.509<br />

5.00 1.668 1.569 1.528<br />

6.00 1.685 1.585 1.544<br />

8.00 1.713 1.611 1.569<br />

10.00 1.735 1.632 1.589<br />

20.00 1.805 1.698 1.654<br />

30.00 1.847 1.737 1.692<br />

40.00 1.877 1.766 1.720<br />

50.00 1.901 1.789 1.742<br />

fDV CwVt � �<br />

2gDi V<br />

where<br />

�Hf = head loss due to friction (in units of length)<br />

fD = Darcy–Weisbach friction factor<br />

C1 = constant<br />

Equation 4-9 may also be reexpressed as<br />

�H f = L� + C 1 � (4-9)<br />

�H f g<br />

� L<br />

4.15<br />

fDV C1CwVt g<br />

= + � (4-10)<br />

V<br />

2<br />

�<br />

2Di


4.16 CHAPTER FOUR<br />

By differentiating this equation with res<strong>pe</strong>ct to V, we obtain for the minimal value<br />

or<br />

2 f DV<br />

� 2Di<br />

f DV<br />

� Di<br />

=<br />

=<br />

–C 1C wV t g<br />

��<br />

V 2<br />

C 1C wV t g<br />

� V 2<br />

V 3 =<br />

at constant friction factor fD, or<br />

[C1CwVt gDi] Vmin = (4-11)<br />

1/3<br />

C1CwVt gDi ��<br />

fD<br />

��<br />

f D 1/3<br />

The magnitude of the Darcy friction factor for water flow in rubber lined and HDPE<br />

pi<strong>pe</strong> was computed for pi<strong>pe</strong>s from 2� to 18� and results presented in Chapter 2.<br />

Wilson (1942) defined a factor C3 to determine whether the particles will settle to<br />

form a bed:<br />

C3 = (4-12)<br />

If C3 > 1 most particles with a terminal velocity Vt will stay in sus<strong>pe</strong>nsion. If C3 � 1 most<br />

particles with a terminal velocity Vt will settle out.<br />

Whereas the equations of Newitt et al. (1955) and Wilson (1942) focused on the terminal<br />

velocity, the work of Durand and Condolios (1952) focused on the drag coefficient for<br />

sand and gravel. Zandi and Govatos (1967) and Zandi (1971) extended the work of Durand<br />

to other solids and to different mixtures. They defined an index number as<br />

V<br />

Ne = (4-13)<br />

At the critical value when Ne = 40, the flow transition between saltation and heterogeneous<br />

regimes occurs. This statement infers that when Ne < 40 saltation occurs, and when<br />

Ne � 40 heterogeneous flow develops. These results, based on a mixture of different particle<br />

sizes, did not apply to the work of Blatch (1906), who concentrated on particles of a<br />

uniform size (sand 20–30 mesh in water). Babcock (1967) reinterpreted this work and<br />

demonstrated that for finely graded particles the transition occurred at an index number of<br />

10. It is obvious that a complex mixture of particles of different sizes can increase the<br />

magnitude of the transition index number.<br />

2 2Vt ��<br />

(�Hf fDgDi/L) 1/2 CD ��<br />

CvDig(�s/�w – 1)<br />

1/2<br />

Example 4-4<br />

Tailings from a mine consist of solids at a volumetric concentration of 20%. The s<strong>pe</strong>cific<br />

weight of the solids is 4.2. The pi<strong>pe</strong> diameter is 8� with a wall thickness of 0.375� and<br />

rubber lining of 0.5�. The particle Albertson sha<strong>pe</strong> factor is 0.7. The dynamic viscosity is<br />

3 cP. The average d50 = 0.4 mm. Determine the s<strong>pe</strong>ed of transition from saltation using<br />

the Zandi approach as expressed by Equation 4-13.<br />

Solution in SI Units<br />

Pi<strong>pe</strong> inner diameter Di = 8� – 2 · (0.5 + 0.375) = 6.25� = 158.75 mm


Iteration 1<br />

Let us first assume a transition from saltation at 3 m/s and let us determine the drag coefficient<br />

of the particles in water at the stated dynamic viscosity:<br />

Rep = 1000 · 0.0004 · 3/0.003 = 400<br />

From Table 3.7, CD = 1.09.<br />

The transition from saltation occurs when Ne = 40. From Equation 4-13, using SI units:<br />

Ne =<br />

Ne = 9.43.<br />

Iteration 2<br />

Let us first assume a transition from saltation at 6 m/s and let us determine the drag coefficient<br />

of the particles in water at the stated dynamic viscosity:<br />

Rep = 1000 · 0.0004 · 6/0.003 = 800<br />

From Table 3.7, CD = 1.15.<br />

Ne =<br />

Ne = 39.<br />

The transition from saltation therefore occurs at a s<strong>pe</strong>ed of 6.1 m/s.<br />

Solution in USCS Units<br />

Iteration 1<br />

Pi<strong>pe</strong> diameter = 8� – 2 · (0.375 + 0.5) = 6.25� = 0.521 ft<br />

Let us first assume a transition from saltation at 10 ft/s and let us determine the drag coefficient<br />

of the particles in water at the stated dynamic viscosity.<br />

Particle size = 0.4 mm/304.7 mm = 1.3128 × 10 –3 ft<br />

� = 0.003/47.88 = 6.265 × 10 –5 lbf-sec/ft2 Density of water = 62.3 lbm/ft3 /32.2 ft/sec = 1.935 slugs/ft3 1.935 slugs/ft<br />

Re =<br />

From Table 3.7, CD = 1.09.<br />

3 × 1.3128 × 10 –3 ft × 10 ft/sec<br />

�����<br />

6.265 × 10 –5 lbf-sec/ft2 9 · �1�.0�9�<br />

����<br />

0.2 · 0.15875 · 9.81 · (4.2/1 – 1)<br />

36 · �1�.1�5�<br />

����<br />

0.2 · 0.15875 · 9.81 · (4.2/1 – 1)<br />

= 406<br />

N e = 9.73.<br />

HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

N e =<br />

100 · �1�.0�9�<br />

����<br />

0.2 · 0.5208 ft · 32.2 ft/sec · (4.2/1 – 1)<br />

4.17<br />

Iteration 2<br />

Let us first assume a transition from saltation at 20 ft/s and let us determine the drag coefficient<br />

of the particles in water at the stated dynamic viscosity:<br />

Rep = 406 · (20/10) = 804


4.18 CHAPTER FOUR<br />

From Table 3.7, CD = 1.15.<br />

20<br />

Ne =<br />

Ne = 39.97.<br />

The transition from saltation therefore occurs at a s<strong>pe</strong>ed of 20 ft/sec.<br />

2 · �1�.1�5�<br />

����<br />

0.2 · 0.5208 ft · 32.2 ft/sec · (4.2/1 – 1)<br />

4-3-3 V4: Transition S<strong>pe</strong>ed Between Heterogeneous and<br />

Pseudohomogeneous Flow<br />

For the transition to pseudohomogeneous flows, Newitt et al. (1955) expressed the s<strong>pe</strong>ed<br />

in terms of the terminal velocity of particles as<br />

V4 = (1800 gDiVt) 1/3 (4-14)<br />

Refer to Chapter 3 and Equation 3-18 to calculate terminal velocity.<br />

Govier and Aziz (1972) applied Newton’s law (i.e., CD = 0.44) for particles immersed<br />

in a fluid to Equation 4-14 to yield<br />

1/3 V4 = 38.7D i (S – 1) 1/6 4gdp � (4-15)<br />

3CD<br />

Govier and Aziz (1972) analyzed the work of S<strong>pe</strong>lls (1955) on solid particles with a<br />

diameter 80 �m < dp < 800 �m (mesh 180 < dp < 20) and derived the following equation:<br />

0.816 0.633 1.63 V4 = 134CD D i V t (4-16)<br />

This equation was derived in USCS units with the diameter expressed in feet and the velocity<br />

in feet <strong>pe</strong>r seconds.<br />

Example 4-5<br />

An ore with a s<strong>pe</strong>cific gravity of 4.1 is to be pum<strong>pe</strong>d in a pseudohomogeneous regime in<br />

a 24 in pi<strong>pe</strong> with an ID of 22.23 in. The drag coefficient of the particles is assumed to be<br />

0.44. The estimated flow rate is 12,000 US gpm. The particles have a sphericity of 0.72<br />

and a diameter of 250 �m. Solve for V4. Solution in SI Units<br />

Q = = 0.757 m3/s<br />

Pi<strong>pe</strong> ID = 22.25 × 0.0254 = 0.565 m<br />

Cross-sectional area = 0.251 m2 Average s<strong>pe</strong>ed of flow = 3.02 m/s<br />

Sphericity = Asp/Ap = 0.72<br />

dsp = �0�.7�2� � 2�5�0� = 218 �m<br />

4 × 0.218 × 10<br />

Vt = ��������<br />

Vt = 0.142 m/s<br />

–3 12,000 × 3.785<br />

��<br />

60,000<br />

× 9.81 (4100 – 1000)<br />

����<br />

3 × 0.44 × 1000


HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

By Newitt’s equation (Equation 4.14):<br />

V4 = (1800 × 9.81 × 0.565 × 0.142) 1/3<br />

V4 = 11.22 m/s<br />

Alternatively using Equation 4.16:<br />

Di = 1.854 ft<br />

Vt = 0.466 ft/sec<br />

0.816 0.633 1.63<br />

V4 = 134C D D i V t<br />

V4 = 134 × 0.440.816 × 1.8540.633 × 0.4661.63 = 29.19 ft/sec or 8.9 m/s<br />

Solution in USCS Units<br />

Q = 12,000 · 0.002228 = 26.736 ft3 /sec<br />

Pi<strong>pe</strong> ID = 22.25/12 = 1.854 ft<br />

Cross-sectional area = 2.7 ft2 Average s<strong>pe</strong>ed of flow = 9.9 ft/sec<br />

Sphericity = Asp/Ap = 0.72<br />

dsp = �0�.7�2� � 2�5�0� = 218 �m = 0.000715 ft<br />

The density of water is 1.93 slugs/ft3 The density of solids is 7.913 slugs/ft3 Vt = �������<br />

Vt = 0.465 ft/s<br />

By Newitt’s equation (Equation 4.14):<br />

V4 = (1800 × 32.2 × 1.854 × 0.465) 1/3<br />

4 × 0.000715 × 32.2 (7.913 – 1.93)<br />

����<br />

3 × 0.44 × 1.93<br />

V4 = 36.83 ft/sec<br />

Alternatively, using Equation 4.16:<br />

0.816 0.633 1.63<br />

V4 = 134C D Di V t<br />

V 4 = 134 × 0.44 0.816 × 1.854 0.633 × 0.466 1.63 = 29.19 ft/sec<br />

4-4 HYDRAULIC FRICTION GRADIENT OF<br />

HORIZONTAL HETEROGENEOUS FLOWS<br />

4.19<br />

Having been able to determine the s<strong>pe</strong>ed for transition from one regime to another, the<br />

<strong>slurry</strong> engineer must determine the loss of head <strong>pe</strong>r unit length due to friction, called the<br />

hydraulic friction gradient (Equation 2-24). The hydraulic friction gradient for the <strong>slurry</strong><br />

(i m) is higher than the hydraulic friction gradient for an equivalent volume of water. Since<br />

the first <strong>slurry</strong> pi<strong>pe</strong>lines were built, engineers and scientists have tried to correlate the<br />

losses with <strong>slurry</strong> to those of an equivalent volume of water.<br />

It was initially assumed that the friction losses would increase in proportion to the vol-


4.20 CHAPTER FOUR<br />

umetric concentration of solids. A term im was then defined as the friction head of the mixture<br />

in equivalent meters (or feet) of the carrier fluid (e.g., water) <strong>pe</strong>r unit of pi<strong>pe</strong> length.<br />

In Chapter 2, the friction hydraulic gradient was introduced by Equation 2-24 and is<br />

defined as:<br />

fDV i =<br />

There are a number of models to predict friction losses and they are essentially based<br />

on the interaction forces between solids and liquid carrier. Some use the drag coefficient,<br />

others use the terminal velocity of the solids, and some consider the solids to be moving<br />

as a bed with a layer of liquid and sus<strong>pe</strong>nded fines above it.<br />

To reflect the increase in friction head due to the volumetric concentration of solids,<br />

Durand and Condolios (1952) proposed a nondimensional ratio<br />

im – iL Z = � (4-17)<br />

CviL where<br />

Cv = the volumetric concentration of solids<br />

im = pressure gradient for the <strong>slurry</strong> mixture in meters of water<br />

iL = pressure gradient for an equivalent volume of water or carrier fluid in meters of water<br />

The reader should not confuse head of <strong>slurry</strong> in meters or feet of <strong>slurry</strong> with meters or<br />

feet of water. This is not a barometer or some instrument measuring pressure; for this reason<br />

everything is kept consistent by using meters or feet of water. By itself, the term i relates<br />

only to clear water having the same velocity as the <strong>slurry</strong> flow. It is convenient to<br />

use water as a reference benchmark. (See Figure 4-9.)<br />

2<br />

�<br />

2gDi<br />

Head loss <strong>pe</strong>r pi<strong>pe</strong> length<br />

in equivalent (m/m) or (ft/ft)<br />

C V2<br />

water<br />

C V1<br />

C<br />

V3<br />

Average velocity of flow<br />

i m<br />

i L<br />

FIGURE 4-9 Concepts of the hydraulic friction gradients i m and i L for <strong>slurry</strong> mixture and for<br />

water.


HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

4.4.1 Methods Based on the Drag Coefficient of Particles<br />

Based on their analysis of test data from 11 references for sand in particle sizes ranging<br />

up to 1 inch (25.4 mm), in pi<strong>pe</strong>s with a diameter range from 1.5 inch to 22 inch, and in<br />

volumetric concentration up to 22%, Zandi and Govatos (1967) derived an equation for<br />

the index number Ne (equation 4-13) in terms of the volumetric concentration, and some<br />

empirical parameters:<br />

� = (4-18)<br />

Or from equation 4-13:<br />

Ne = (4-19)<br />

or � = CvNe. They plotted this function against a parameter � to express head loss as<br />

� = = K(�) m (4-20)<br />

where<br />

iL = hydraulic gradient in terms of water density for a flow of clean water with a mean<br />

velocity V<br />

im = hydraulic gradient in terms of water density for a <strong>slurry</strong> flow with a mean velocity<br />

V<br />

K, m = constants<br />

On a logarithmic scale they obtained:<br />

For � > 10, K = 6.3 and m = –0.354<br />

For � < 10, K = 280 and m = –1.93<br />

The data is presented in Figure 4-10.<br />

The dramatic change in values of K and m at � = 10 has encouraged researchers to develop<br />

more sophisticated models that we shall review in the rest of this chapter.<br />

Substituting for the value of 40 of the index coefficient, V3 may be expressed as<br />

[40 CvDi g(�s – �w)/�w] V3 = (4-21)<br />

1/2<br />

V<br />

�<br />

�<br />

Cv<br />

im – iL �<br />

CviL ���<br />

2 1/2 C D<br />

��<br />

Dig(�s/�w – 1)<br />

C D 1/4<br />

4.21<br />

Equation 4-21 is therefore a modified version of Equation 4-2. Equation 4-4 is a different<br />

approach, as it accounts for particle size, which is often easier to measure than the<br />

drag coefficient. Example 4-4 has shown that some iteration is necessary to obtain the velocity<br />

at which the transition from saltation to asymmetric flow occurs.<br />

Despite its simplicity, this method continues to be used by dredging engineers who<br />

usually deal with sand and gravel mixtures of less than 20% concentration by volume.<br />

The <strong>pe</strong>rsonal ex<strong>pe</strong>rience of the author is that often mines and dredging <strong>systems</strong> have<br />

to be designed in very remote areas where there are no <strong>slurry</strong> labs to conduct tests. This is<br />

an unfortunate fact, and sometimes an “overconservative” approach based on Durand,<br />

Zandi, and other authors is the only alternative. However, the author does encourage engineers<br />

of <strong>slurry</strong> <strong>systems</strong> to plan well ahead and test data to avoid very ex<strong>pe</strong>nsive field corrections.


4.22 CHAPTER FOUR<br />

i L<br />

� Cv i L<br />

� =<br />

10000<br />

1000<br />

100<br />

10<br />

1<br />

.1<br />

Durand<br />

&<br />

Condolios<br />

FIGURE 4-10 The Zandi–Govatos factors for heterogeneous <strong>slurry</strong> flows. (From Zandi and<br />

Govatos, 1967, reprinted with <strong>pe</strong>rmission from ASCE.)<br />

Shook et al. (1981) modified Zandi’s equation by proposing “in-situ concentration of<br />

particles” C t rather than volumetric concentration:<br />

� t = K� m<br />

� t =<br />

i m – i L<br />

� iLC t<br />

Zandi &<br />

Govtes<br />

RANGE OF<br />

1 NUMBER<br />

� 0–40<br />

� 40–310<br />

� 310–1550<br />

� 1550–3100<br />

.01<br />

.01 0.1 1.0 10 100 1000<br />

V<br />

� =<br />

2�C� D�<br />

��<br />

Di g(�s/�L – 1)<br />

They measured a magnitude of m = –1 for one single ty<strong>pe</strong> of coal in different pi<strong>pe</strong> sizes.<br />

They measured different values of K for different coals. The in-situ concentration C t remained<br />

constant with s<strong>pe</strong>ed, but the volumetric concentration of solids C v that could be<br />

moved increased with V. This concept will be reexamined in Section 4.10 as part of the<br />

two-layer models.<br />

Example 4-6<br />

Using Equations 4-19 to 4-20, consider the pumping of solids in a 305 mm (12 in) ID pi<strong>pe</strong><br />

at a s<strong>pe</strong>ed of 3.045 m/s (10 ft/s) and a volumetric concentration of 18%. Assume a drag<br />

coefficient of 0.45 for the solid particles and a s<strong>pe</strong>cific gravity of 2.65. Determine the increase<br />

in the pressure gradient for flow in the pi<strong>pe</strong> due to the presence of solids.


Solution in SI Units<br />

V = 3.045 m/s<br />

pi<strong>pe</strong> Di = 0.305 m (12 in)<br />

3.045<br />

Ne = = 68.67<br />

Ne 68.67<br />

� = � = � = 381.5<br />

Cv 0.18<br />

� > 10 then K = 6.3 and m = –0.345<br />

2 × �0�.4�5�<br />

���<br />

0.18 × 0.305 × (2.65 – 1)<br />

� = = K � –0.345 = 0.81<br />

= 0.81 × 0.18 = 0.145<br />

im � = 1.145<br />

iL<br />

The <strong>slurry</strong> causes an increase of pressure gradient of 14.5% by comparison with water<br />

at the same velocity.<br />

Using the approach develo<strong>pe</strong>d by Durand and Condolios, the fanning friction factor<br />

for the <strong>slurry</strong> is correlated with the friction factor for an equivalent volume of water by<br />

the following equation:<br />

gDi(�s – �L) 3/2<br />

fDm = fDL�1 + Kf Cv���� �<br />

(4-22)<br />

2 V �L�C� D�<br />

Wasp et al. (1977) deducted that the coefficient Kf is between 80 and 150, de<strong>pe</strong>nding<br />

on the <strong>slurry</strong>. The most common value is actually 81 for most sands according to Govier<br />

and Aziz (1972) (see Table 4-5).<br />

Example 4-7<br />

Using Equation 4-22, determine the correction for the friction factor for the portion of<br />

solids in a <strong>slurry</strong> mixture of uniform size distribution. The <strong>slurry</strong> is pum<strong>pe</strong>d at the rate of<br />

16,000 gpm in a rubber-lined 22.75� ID pi<strong>pe</strong>. The volumetric concentration is 22%. Assume<br />

K f = 85 and C D = 0.45. Use the Swain–Jaime equation to determine f L. The s<strong>pe</strong>cific<br />

gravity of the solids is 2.65. The dynamic viscosity of water is 2.7 × 10 –5 lbf-sec/ft 2 .<br />

Solution in SI Units<br />

HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

i m – i L<br />

� iL<br />

Q = = 1.009 m3 /s<br />

Pi<strong>pe</strong> ID = 22.75 (0.0254) = 0.5778 m<br />

Area of pi<strong>pe</strong> = 0.262 m2 16,000 (3.785)<br />

��<br />

60,000<br />

Velocity = 3.85 m/s<br />

Dynamic viscosity = 0.00129 mPa · s<br />

4.23


4.24 CHAPTER FOUR<br />

TABLE 4-5 Correction of Friction Factor Due to Volumetric Concentration of Solids<br />

Based on Equation 4-22 Assuming K = 81<br />

gD i (� s – � L)<br />

��<br />

V 2 �L�C� D�<br />

For the water:<br />

1,000 (3.85) 0.5778<br />

Re = ��� = 1,723,292<br />

0.00129<br />

Absolute roughness of rubber = 0.00015 m.<br />

Relative roughness<br />

� 0.00015<br />

� = � = 0.0002596<br />

DI 0.5778<br />

0.25<br />

fD = ����� = 0.0151<br />

[log10{(0.0002596/3.7) + (5.74/1,723,2920.9 )} 2 ]<br />

Solution in USCS Units<br />

f Dm – f DL<br />

� CV f DL<br />

0.01 0.081 0.45 24.451<br />

0.02 0.229 0.50 28.638<br />

0.03 0.421 0.55 33.039<br />

0.04 0.648 0.60 37.645<br />

0.05 0.906 0.65 42.448<br />

0.06 1.190 0.70 48.024<br />

0.07 1.500 0.75 52.611<br />

0.08 1.833 0.80 57.959<br />

0.09 2.187 0.85 63.477<br />

0.10 2.561 0.90 69.159<br />

0.15 4.706 0.95 75.002<br />

0.20 7.245 1.00 81.000<br />

0.25 10.125<br />

0.30 13.31<br />

0.35 16.77<br />

0.40 20.49<br />

fm = fL�1 + 85 · 0.22� � 1.5<br />

9.81 · 0.578 · 1.65<br />

�� �<br />

f m = f L · 18.067 = 0.273<br />

Q = 35.63 ft3 /sec<br />

22.75<br />

Pi<strong>pe</strong> ID = � = 1.896 ft<br />

12<br />

Area = 2.823 ft 2<br />

gD i (� s – � L)<br />

��<br />

V 2 �L�C� D�<br />

3.85 2 �0�.4�4�5�<br />

f Dm – f DL<br />

� CV f DL


Velocity = 12.62 ft/s<br />

Dynamic viscosity = 1.29 cP = 0.0129 · 0.002089 lbf-sec/ft2 = 0.00002695 lbf-sec/ft2 Re = � � = 1.7 × 106 Absolute roughness of rubber = 0.00049 ft<br />

Relative roughness of rubber = 0.0002596<br />

fD = 0.0151<br />

fm = 0.0151�1 + 85 × 0.22� � 1/5<br />

32.2 × 1.896 × 1.65<br />

���� � = 0.27<br />

12.622 62.3 12.62 (1.896)<br />

� ��–4<br />

32.2 2.695 × 10<br />

�0�.4�5�<br />

An increase of the friction factor by 18-fold ap<strong>pe</strong>ars to be very high. The engineer in<br />

charge of such a problem should seriously consider redesigning the system. At this stage,<br />

the reader is encouraged to become familiar with the basic equations before applying<br />

them to compound <strong>systems</strong>.<br />

Equation 4-17 can be expressed in terms of the drag coefficients of the solid particles,<br />

the pi<strong>pe</strong> inner diameter, the density of the solid and liquid phases, the s<strong>pe</strong>ed, and an ex<strong>pe</strong>rimental<br />

factor Ke: im – iL Dig(�s/�L – 1) 1 3/2<br />

Z = � = Ke������� (4-23)<br />

2<br />

CviL V �C�D�<br />

Babcock (1968) was very critical of all equations using pressure gradients based on<br />

the work of Durand and Condolios or their followers. Geller and Gray (1986) did not<br />

agree with Babock’s criticisms and s<strong>pe</strong>lled out some of the misgivings. Govier and Aziz<br />

(1972) did confirm that errors of the order on 40% have occurred in predicted values of Z,<br />

but for all intents and purposes, these equations were the best available till the early<br />

1970s. Herbich (1991) agreed with the value of 81 for most dredged sands and gravel.<br />

Sand and gravel are typically dredged, then pum<strong>pe</strong>d at a volumetric concentration smaller<br />

than 20%.<br />

4.4.2 Effect of Lift Forces<br />

HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

4.25<br />

It may be considered that the magnitude of the constant m is based on a very large magnitude<br />

of data. In an innovative study at the Canada Center for Mineral and Energy Technology<br />

(CANMET), Geller and Gray (1986) conducted an extended analysis that demonstrated<br />

that lift forces had an effect on the pressure gradient. This study, rather than<br />

dismissing the ideas of Durand, supported the previous work and gave it more importance.<br />

Reviewing the work of Babock (1971), Geller and Gray (1986) indicated that for fine<br />

to intermediate sizes (80/100 quartz sand with d � = 0.16 mm) the value of m was –0.25. In<br />

addition, they concluded that lift forces are at a maximum when the volumetric concentration<br />

C v is less than 0.23. For intermediate sands at higher volumetric concentration, the<br />

lift forces seem to be minimal. This is an important factor to consider (for an understanding<br />

of lift forces review Chapter 3, Section 3.1).<br />

Furthermore, there is an important coefficient of mechanical friction � p, which results<br />

from the sliding displacement between solids in contact, which is distinct from the viscous<br />

friction.


4.26 CHAPTER FOUR<br />

4-4-3 Russian Work on Coarse Coal<br />

There are no universally accepted models for coarse coal. Work in the former Soviet<br />

Union on coarse coal was reported by Traynis (1970) and reviewed by Faddick (1982).<br />

From Russian data, the following two equations were reported. For deposition velocity:<br />

V3 = [Dig] 1/2 [(�c – �hm)/�c] (4-24)<br />

For the hydraulic gradient for coal:<br />

1/3<br />

���<br />

[ fDLkCD] 1/3<br />

�s – �L Cvc(�s – �hm) � � ��<br />

�L k CdV�L where<br />

Cv = total volumetric concentration of solids<br />

Cvc = volumetric concentration of coarse solids<br />

K = constant for coarse coal = 1.9<br />

CD = drag coefficient considered to be 0.75 for the coarse coal fraction<br />

�hm = density of heavy medium produced by the fines<br />

i m = i L� 1 + C v� � + � · �� (4-25)<br />

For the other terms, see Section 4-14.<br />

�g�D� i�<br />

Example 4-8<br />

Coarse coal is to be pum<strong>pe</strong>d in a rubber-lined 18 in pi<strong>pe</strong> steel with an inner diameter of 17<br />

in. A screen analysis of the coal indicates that it has a distribution of 20% passing 200 microns.<br />

The velocity of pumping is 4.5 m/s and the total weight concentration is 52%. The<br />

s<strong>pe</strong>cific gravity of the coal is 1.35. Determine the hydraulic gradient due to wall friction<br />

in the horizontal pi<strong>pe</strong>line. Assume a water dynamic viscosity of 1.2 cP, but correct for<br />

viscosity due to solids using Einstein’s equation. Assume a drag coefficient of 0.75 for<br />

the coarse coal.<br />

Solution<br />

Since the weight concentration is 52%, the s<strong>pe</strong>cific gravity of the mixture is<br />

Sm = SL/(1 – (CW(Ss– SL)/Ss) = 1/(1 – 0.52(1.35 – 1)/1.35) = 1.156<br />

The volumetric concentration is<br />

Cv = CwSm/Ss = 0.52 · 1.156/1.35 = 0.445<br />

The weight concentration of the fines is 20%.<br />

Density of the heavy medium carrying the fines is<br />

Smf = SL/(1 – (CWf(Ss– SL)/Ss) = 1/(1 – 0.104(1.35 – 1)/1.35) = 1.028<br />

Volumetric concentration of the fines = 0.2 · 0.445 = 0.089.<br />

Calculations in SI Units<br />

Pi<strong>pe</strong> ID = 17 (0.0254) = 0.432 m<br />

Area of pi<strong>pe</strong> = 0.146 m2 Velocity = 4.5 m/s<br />

The dynamic viscosity is corrected to take in account the presence of fines at a volumetric<br />

concentration of 0.089. The dynamic viscosity of water is 1.2 cP, the Einstein–Thomas<br />

equation is applied:


HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

� = �L(1 + (2.5 · 0.089) + (10.05 · 0.0892 ) + 0.00273 exp (16.6 · 0.089)] = 1.314<br />

�L = 1.577cP<br />

1,000(4.5) 0.432<br />

Re = �� = 1,232,720<br />

0.001577<br />

Absolute roughness of rubber = 0.00015 m.<br />

Relative roughness:<br />

� 0.00015<br />

� = � = 0.000368<br />

DI 0.432<br />

fDL = = 0.0162<br />

iL = fDV 2 /(2gDi) = 0.0162 · 4.52 0.25<br />

�����<br />

[log10{(0.000368/3.7) + (5.74/1,232,720<br />

/(2 · 9.81 · 0.432) = 0.0387 m/m<br />

Using Equation 4.25:<br />

0.9 )} 2 ]<br />

im = �1 + 0.445� � + 1350 – 1000 �(9�.8�1� ·� 0�.4�3�2�)� 0.8 · 0.445 · (1350 – 1028)<br />

�� � �� · �����<br />

im = 0.0387[1 + 0.445(0.35)] + [0.0368)] = 0.0815m/m<br />

The presence of coal effectively doubles the head losses. The deposition velocity is<br />

expressed from Equation 4-24:<br />

V3 = [0.432 · 9.81] 1/2<br />

[(1350 – 1028)/1350] 1/3<br />

���<br />

[0.0162 · 1.9 · 0.75] 1/3<br />

1000 1.9 · 0.75 · 4.5<br />

1000<br />

V3 = 4.48 m/s<br />

Calculations in USCS Units<br />

Pi<strong>pe</strong> ID = 17� = 1.417 ft<br />

Area of pi<strong>pe</strong> = 1.576 ft2 Velocity = 4.5 m/s = 14.76 ft/sec<br />

The dynamic viscosity is corrected to take in account the presence of fines at a volumetric<br />

concentration of 0.089. For the water, dynamic viscosity = 1.2 cP = 0.012 · 0.002089 lbfsec/ft2<br />

= 2.507 × 10 –5 lbf-sec/ft2 . The Einstein–Thomas equation is applied:<br />

� = � L(1 + (2.5 · 0.089) + (10.05 · 0.089 2 ) + 0.00273 exp (16.6 · 0.089)] = 1.314 � L<br />

= 3.294 × 10 –5 lbf-sec/ft 2 .<br />

For the water, the density is 1.934 slugs/ft3 .<br />

Re = = 1.23 × 106 1.934 · 14.76 · 1.417<br />

���<br />

3.294 × 10 –5<br />

Absolute roughness of rubber = 0.000492 ft.<br />

Relative roughness:<br />

�<br />

� DI<br />

0.000492<br />

= � = 0.000368<br />

1.417<br />

4.27


4.28 CHAPTER FOUR<br />

fDL = = 0.0162<br />

iL = fDV 2 /(2gDi) = 0.0162 · 14.762 0.25<br />

�����<br />

[log10{(0.000368/3.7) + (5.74/(1.23 × 10<br />

/(2 · 32.2 · 1.417) = 0.0387 ft/ft<br />

Using Equation 4.25, and substituting density with s<strong>pe</strong>cific gravity<br />

6 ) 0.9 )} 2 ]<br />

im = �1 + 0.44� � + 1.350 – 1 �(3�2�.2� �· 1�.4�1�7�)� 0.8 · 0.445 · (1.350 – 1.028)<br />

� �� · �����<br />

�<br />

1 1.9 · 0.75 · 14.76<br />

1.0<br />

im = 0.0387[1 + 0.445(0.35)] + [0.0368)] = 0.0815ft/ft<br />

The presence of coal effectively doubles the head losses. The deposition velocity is<br />

expressed from Equation 4-24:<br />

V 3 = [1.417 · 32.2] 1/2<br />

V3 = 14.71 ft/sec<br />

The coal <strong>slurry</strong> is therefore being pum<strong>pe</strong>d just above the deposition s<strong>pe</strong>ed, and therefore<br />

at the minimum pressure gradient for horizontal pi<strong>pe</strong>lines.<br />

4-4-4 Equations for Nickel–Water Sus<strong>pe</strong>nsions<br />

Ellis and Round (1963) conducted tests on a mixture of nickel particles and water and derived<br />

the following equation:<br />

� = = K(�) m = 385 � –1.5 im – iL � (4-26)<br />

CviL The constants K and m are therefore different from those reported by Zandi and Govatos<br />

(1967) for sand particles, as expressed by Equation 4-20.<br />

4-4-5 Models Based on Terminal Velocity<br />

Newitt et al. (1955) conducted tests in pi<strong>pe</strong>s smaller than 150 mm (6 in) and proposed to<br />

express Z in terms of the terminal velocity (instead of the drag coefficient).<br />

Z = = K2� � (4-27)<br />

where<br />

K2 = an ex<strong>pe</strong>rimentally determined constant. For small pi<strong>pe</strong>s, K2 = 1100.<br />

Vm = mean velocity of mixture<br />

For solids of different sizes, Newitt suggested a weighted mean diameter as<br />

dpm = � n<br />

im – i �s – �L gDiVt � � �3 Cvi �L V m<br />

dpimi/mt (4.28)<br />

where<br />

m i = the mass of solids with particle diameter of d p<br />

m t = total mass of solids<br />

[(1.350-1.028)/1.350] 1/3<br />

���<br />

[0.0162 · 1.9 · 0.75] 1/3<br />

i=1


HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

4.29<br />

Hayden and Stelson (1968) proposed a modification of the Durand–Condolios equation<br />

using the terminal velocity instead of the drag coefficient:<br />

= 100� � 1.3<br />

gDi[(�m – �L)/�L]Vt ���<br />

(4-29)<br />

V<br />

Geller and Gray (1986) pointed out that the equations of Durand, Newitt, and Babcock<br />

converged when m = –1.<br />

Newitt et al. (1955) minimized the importance of lift forces when a bed cannot form<br />

because of lift forces on particles. However, the work of Bagnold (1954, 1955, 1957) indicated<br />

that the submerged weight of particles separated from the bed was transmitted to<br />

the bed or the pi<strong>pe</strong> wall under the same conditions. Thus, mechanical friction can contribute<br />

to head loss.<br />

It may be argued that sometimes it is easier to measure the terminal velocity rather<br />

than the drag coefficient, particularly with oddly sha<strong>pe</strong>d particles. As Chapter 3 clearly<br />

demonstrated, both parameters are interrelated.<br />

2 im – iL �<br />

CviL �g�d� ��� p( m�)/ ��� L�–� 1�)�<br />

Example 4-9<br />

The tailings from a small mine are pum<strong>pe</strong>d at a weight concentration of 40%. They consist<br />

of crushed rock at a s<strong>pe</strong>cific gravity of 3.2. The d85 of the particles is 1mm. For a flow rate<br />

of 280 m3 /hr, a smooth high-density polyethylene pi<strong>pe</strong> with an internal diameter of 138 mm<br />

is selected. Using Newitt’s method as expressed By equations 4.27 and 4.29, determine the<br />

head loss due to the presence of solids, assuming a dynamic viscosity of 1.8 cP.<br />

Solution in SI Units<br />

Pi<strong>pe</strong> flow area = 0.25 · � · 0.1382 = 0.01496 m2 Average velocity of flow = Q/A = (280/3600)/0.01496 = 5.2 m/s<br />

Particle Reynolds number using the density of water = Rep = 0.001 · 3.71 · 1000/0.0018<br />

= 2063<br />

Since Rep > 800, the flow is turbulent and Newton’s law is used to calculate the terminal<br />

velocity:<br />

Vt = 1.74(dp · g · (�p – �L)/�L) 1/2 = 1.74(0.001 · 9.81 · 2.1) 1/2 = 0.25 m/s<br />

By Newitt’s method, the transition between saltation and motion occurs at 17Vt or<br />

V3 = 17 · 0.25 = 4.25 m/s<br />

Since the weight concentration is 40%, the s<strong>pe</strong>cific gravity of the mixture is<br />

Sm = SL/(1 – (CW(Ss– SL)/Ss) = 1/(1 – 0.4(3.1 – 1)/3.1) = 1.372<br />

The volumetric concentration is Cv = (1.372 – 1)/2.1 = 0.177<br />

Using equation 4.27, and assuming K2 = 1100,<br />

Z = 1100 · (2.1) · (9.81 · 0.138 · 0.25/5.23 ) = 5.563<br />

im/i = 1 + 0.177 · 5.563 = 1.985<br />

Using equation 4.29:<br />

= 100� � 1.3<br />

9.81 · 0.138 · 2.1 · 0.25<br />

��� = 11<br />

5.2<br />

im/iL = 1 + 0.177 · 11 = 2.95<br />

2 [9.81 · 0.001 · 2.2) 1/2<br />

im – iL �<br />

CviL


4.30 CHAPTER FOUR<br />

This example and the use of these two equations indicates that the empirical coefficients<br />

of 1100 in the Newitt method for fine coal and sand, or the empirical coefficient of 100<br />

for sand from the Hayden and Stelson equation, do not converge for similar results. Testing<br />

would be recommended to confirm the magnitude of these coefficients.<br />

4.5 DISTRIBUTION OF PARTICLE<br />

CONCENTRATION IN COMPOUND SYSTEMS<br />

The reader may be familiar with the concepts develo<strong>pe</strong>d in the 1950s and 1960s on uniformly<br />

graded solid particles. In reality, slurries often consist of a wide distribution of<br />

particles. The coarser ones tend to move at the bottom of the horizontal pi<strong>pe</strong>, and the finer<br />

ones move above these bottom layers. Understanding the distribution of these particles<br />

in layers above layers is essential for a correct estimation of the friction losses.<br />

Initially, the work was done in the 1930s and 1940s on o<strong>pe</strong>n channel flows and is<br />

discussed in Chapter 6, Section 6-2-3. The distribution of volumetric concentration is<br />

shown to be a function of depth of the liquid in an o<strong>pe</strong>n channel flow, raised to a exponent.<br />

The exponent is a function of the relation of the terminal velocity to the friction<br />

velocity.<br />

Ismail (1952) was the first to extend the approach of Vanoni to closed conduits. He focused<br />

initially on rectangular closed conduits. This test work demonstrated that the concentration<br />

was an exponential function:<br />

C Vt Log10� � = (y – a) (4-30)<br />

� �<br />

CA Es<br />

where<br />

Es = the mass transfer coefficient<br />

a = height of layer A above bottom of the conduit<br />

y = distance from the lower boundary<br />

C = volumetric concentration of the particle diameter under consideration<br />

CA = volumetric concentration of height “A”<br />

For many pi<strong>pe</strong>s, C/CA is considered by Wasp et al. (1977) to be 0.08 DI from the top of<br />

the pi<strong>pe</strong>. Wasp et al. (1977) examined the distribution of concentration of The Consolidation<br />

Coal Company’s Ohio coal pi<strong>pe</strong>line at a height of 8% from the bottom of the conduit<br />

and at 8% from the top of the conduit; they reinterpreted the work of Ismail (1951) and<br />

devised the following equation:<br />

C 1.8 Vt log10 � = –��� (4-31)<br />

CA �KxUf where<br />

Uf is the friction velocity (discussed in Chapter 2)<br />

Kx is the Von Karman constant<br />

� = constant of proportionality<br />

Hsu et al. (1971) reexamined the work of Ismail by proposing a polar coordinate system<br />

(r, �) for the analysis of the distribution of concentration in a pi<strong>pe</strong>:<br />

C(r, �)<br />

� C(0, 0)<br />

V t<br />

� Uf<br />

r cos � cos �<br />

� ��<br />

RI me<br />

= exp� � �� (4-32)


HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

where<br />

� = the angle from the horizontal<br />

� = angle from the vertical starting at the lowest quadrant point<br />

RI = inner diameter of pi<strong>pe</strong><br />

r = local radius for a point in the flow<br />

Equation 4-30 can be reduced to<br />

C Vt log10 = � �<br />

� � (constant) (4-33)<br />

CA Uf<br />

The extent by which the Von Karman constant Kx is suppressed by turbulence is difficult<br />

to assess.<br />

Ip<strong>pe</strong>n (1971) conducted an analysis of turbulent sus<strong>pe</strong>nsions in o<strong>pe</strong>n channel flows.<br />

This work showed that the concentration close to the lower boundary was the most important<br />

factor suppressing the Von Karman constant. This may not be astonishing when we<br />

consider that beds of coarse particles form in this region at low s<strong>pe</strong>eds.<br />

Hunt (1969) develo<strong>pe</strong>d an equation for diffusion of heterogeneous flows:<br />

d(Cv) ES � + (1 – Cv)CvVt = 0 (4-34)<br />

d(y)<br />

where Cv is the volumetric concentration of solids.<br />

This equation shows that when coarse and fine particles are pum<strong>pe</strong>d together under<br />

certain conditions, the flows may exhibit an increase in concentration of fine particles<br />

with increasing height.<br />

Example 4-10<br />

Using Hunt’s equation, prove that the ratio of concentration at 0.08 DI from the top is the<br />

concentration at pi<strong>pe</strong> center expressed by<br />

VR<br />

log10��� = –1.8 Z<br />

VRa<br />

where<br />

VR = Cv/1 – Cv a = the reference plane at 0.08 DI It has already been shown in Equation (4-31) that<br />

C Vt log10��� = –1.8�<br />

CA �KxUf Let us confirm that Hunt’s approach applies:<br />

E s<br />

dC v<br />

� dy<br />

+ (1 – C v)C vV t = 0<br />

VR =<br />

DC v<br />

� dVR<br />

C v<br />

� 1 – Cv<br />

= (1 – C v) 2<br />

4.31


4.32 CHAPTER FOUR<br />

But some of Hunt’s equation shows that<br />

Then<br />

Or<br />

–V t<br />

� Es<br />

=<br />

(1 – Cv) Cv = · = (1 – Cv) 2<br />

dC dVR<br />

dV<br />

� � �<br />

dVR dy<br />

dy<br />

E s<br />

dC<br />

� dy<br />

–V t<br />

� Es<br />

This is the same as the Equation 4-34.<br />

= � � (1 – Cv) 2<br />

dC dC dVR dVR<br />

� � � �<br />

dy dVR dy<br />

dy<br />

–V t(1 – C v)C v<br />

��<br />

Es<br />

dVR<br />

Cv = � (1 – C<br />

dy<br />

v)<br />

dVR<br />

� (1 – Cv) + VtCv = 0<br />

dy<br />

dVR Cv Es � + Vt � = 0<br />

dy (1 – Cv)<br />

The approach discussed in the previous paragraph is sometimes classified as the distributed<br />

concentration approach. The analysis is based on establishing the plane for reference<br />

CA, usually at 0.08 diameter.<br />

It has been demonstrated that<br />

C<br />

Vt log10��� = – 1.8�<br />

CA �KxUf If � is assumed to be unity and there is no suppression for the Von Karman constant,<br />

i.e., Kx = 0.4, then<br />

C Vt log10��� = –4.5��� (4-35)<br />

CA Uf<br />

Thomas (1962) commented that the Durand–Condolios approach was limited to sand<br />

and similar solids and proposed a more general criterion of evaluating flow of slurries in<br />

terms of the ratio Vt/Uf or ratio of free-fall velocity to friction velocity. He indicated that<br />

when<br />

Vt � > 0.2 (4-36)<br />

Uf<br />

the solids would be transported as a heterogeneous <strong>slurry</strong>.<br />

Charles and Stevens (1972) suggested that Equation 4.32 should be modified to correspond<br />

to C/CA < 0.13, whereas Charles and Stevens’ criterion corresponds to C/CA < 0.27.<br />

The Thomas criterion as expressed by Equation 4-31, corresponds to C/CA < 0.13,<br />

whereas the Charles and Stevens’ criterion corresponds to C/CA < 0.27.


HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

Thomas (1962) indicated that the minimum transport condition for particles de<strong>pe</strong>nds<br />

on a number of factors, and derived the following equation for glass beads:<br />

= 4.90� �� � 0.60<br />

� � 0.23<br />

Vt dpUf 0 � �S – �L � � � � (4-37)<br />

Uf � DiUf 0 �L<br />

where<br />

� = kinematic viscosity of water<br />

Uf 0 = friction velocity at deposition for limiting case of infinite dilution<br />

Thomas (1962) defined a critical friction velocity at which the <strong>slurry</strong> starts to deposit<br />

for a given concentration as<br />

UfC= Uf 0�1 + 2.8 (�C� V�)� � 1/3 Vt �<br />

(4-38)<br />

The approach of Thomas is implicit. It means that to predict U f, it is important to<br />

measure friction loss as a function of velocity. It is then necessary to establish the deposition<br />

velocity using Equations 4-34, 4-35, and 4-36.<br />

4-6 FRICTION LOSSES FOR COMPOUND<br />

MIXTURES IN HORIZONTAL<br />

HETEROGENEOUS FLOWS<br />

Many slurries resulting from dredging, cyclone underflow, and tailings disposal are not<br />

pum<strong>pe</strong>d with single-sized particles. Some authors such as Newitt et al. (1955) proposed<br />

the use of a weighted average particle diameter but Hill et al. (1986) proposed that the<br />

particles should be divided. The finer particles would move as a heterogeneous flow,<br />

while the coarser particles would move as a bed by saltation. The equations of friction<br />

loss for each fraction or size of solids should be calculated as in Sections 4-4-1 and 4-4-3.<br />

Hill et al. (1986), Wasp et al. (1977), and Gaesler (1967) demonstrated that this approach<br />

worked well when applied to pumping water–coal mixtures.<br />

The compound or heterogeneous–homogeneous system is the most important and<br />

most common in <strong>slurry</strong> transportation. It involves coarse and fine particles. The fines<br />

move as a homogeneous mixture while the remainder move as a heterogeneous mixture.<br />

To conduct this analysis, the rheological and physical pro<strong>pe</strong>rties of the solids must be<br />

known.<br />

This method was pioneered by Wasp et al. (1977) and in some res<strong>pe</strong>cts was further develo<strong>pe</strong>d<br />

by the “stratification model” described later on. The heterogeneous mixture or<br />

bed motion is based on the method of concentration in relation to a reference layer, as described<br />

by Equation 4-30.<br />

The method proposed by Wasp et al. (1977) can be summarized as follows:<br />

1. Divide the total size fraction into a homogeneous fraction using Durand’s equation.<br />

2. Calculate the friction losses of the homogeneous fraction based on the rheology of the<br />

<strong>slurry</strong>, assuming Newtonian flow.<br />

3. Calculate the friction losses of the heterogeneous fraction using Durand’s equation.<br />

4. Define a ratio C/C A for the size fraction of solids based on friction losses estimated in<br />

steps 2 and 3.<br />

� Uf 0<br />

4.33


4.34 CHAPTER FOUR<br />

5. Based on the value of C/C A, determine the fraction size of solids in homogeneous and<br />

heterogeneous flows.<br />

Re-iterate steps 2 to 5 until convergence of the friction loss.<br />

Example 4-11<br />

A nickel ore <strong>slurry</strong> needs to flow by gravity at a weight concentration of 28%. The design<br />

flow rate is 1631 m3 /hr. The <strong>slurry</strong> was tested in a 159 mm pi<strong>pe</strong>line with a roughness coefficient<br />

of 0.016 mm at a weight concentration of 26.3%. The results of the pressure drop<br />

versus s<strong>pe</strong>ed are presented in Table 4-2. No data was made available on the drag coefficients<br />

or terminal velocity of the solids.<br />

The particle size distribution of the originally milled ore is presented in Table 4-3.<br />

S<strong>pe</strong>cial screens would be installed to screen away the coarsest particles (larger than 0.850<br />

mm). Conducting a friction loss for a rubber lined steel pi<strong>pe</strong> would be a better option.<br />

(See Tables 4-6 and 4-7.)<br />

The solids density was measured as 4074 kg/m3 . At a weight concentration of 26.3%,<br />

this corresponds to a <strong>slurry</strong> density of 1244 kg/m3 . Volumetric concentration is<br />

�m CV = CW = 0.08%<br />

� �s<br />

Using the Thomas–Einstein equation for dynamic viscosity correction:<br />

� = �L(1 + (2.5 · 0.08) + (10.05 · 0.082 ) + 0.00273 exp(16.6 · 0.08)] = 1.274 · �L Analysis of Test Results<br />

Water at a tem<strong>pe</strong>rature of 20° Celsius has a dynamic viscosity of 1 mPa · s. Slurry viscosity<br />

is therefore 1.274 mPa · s, and the Reynolds number is<br />

1244(V)DI Re = �� = 155,256(V) = 294,986<br />

–3<br />

1.274 × 10<br />

where V = 1.9 m/s<br />

The <strong>slurry</strong> was tested in a pumping test loop. The lab tests indicated a pressure drop of<br />

270 Pa/m at this velocity. The +0.850 mm solids were screened away prior to pump tests.<br />

TABLE 4-6 Pressure Drop versus S<strong>pe</strong>ed in a 159 mm ID Steel Pi<strong>pe</strong> at a Weight<br />

Concentration of 26.3% (Example 4-11)<br />

Tem<strong>pe</strong>rature 20°C Tem<strong>pe</strong>rature 35°C<br />

______________________________________ ______________________________________<br />

Velocity (m/s) Pressure drop (kPa) Velocity (m/s) Pressure drop (kPa)<br />

0.61 0.063<br />

1.00 0.085 1.00 0.079<br />

1.5 0.175 1.51 0.169<br />

1.9 0.270 1.91 0.259<br />

2.3 0.360 2.30 0.358<br />

2.7 0.525 2.70 0.487<br />

3.1 0.688 3.11 0.628<br />

3.5 0.847 3.50 0.793<br />

4.0 1.046 4.00 0.988


HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

TABLE 4-7 Particle Size Distribution Prior to Screening the<br />

Coarsest Solids (Example 4-11)<br />

Size (mm) Volumetric concentration<br />

+ 0.850 14.3%<br />

–0.850 to +0.400 1.61%<br />

–0.400 to +0.200 1.91%<br />

–0.200 to +0.105 1.41%<br />

–1.05 to +0.044 1%<br />

–0.044 79.8%<br />

4.35<br />

Table 4-8 indicates the new volumetric concentration of the solids in the <strong>slurry</strong> after<br />

screening the +0.850 mm solids.<br />

The method develo<strong>pe</strong>d by Wasp et al. (1977) has been used very successfully over the<br />

last 25 years for Newtonian slurries and will be used in the present calculations. The<br />

roughness of a steel pi<strong>pe</strong> is 0.046 mm. Assuming that the –0.044 mm particles were transported<br />

by turbulence above the moving bed of coarser particles, the Swain–Jain equation<br />

may be used in the range of 5000 < Re < 100,000,000 to determine the friction coefficient<br />

of the homogeneous part of the mixture:<br />

fD = = 0.017<br />

where fD = the Darcy friction factor<br />

For the density of 1244 kg/m3 , the pressure losses of the carrier fluid (including the<br />

–0.044 mm) at a first iteration is therefore<br />

0.017(1.9<br />

Loss = = 240 Pa/m<br />

The lab test measured 270 Pa/m; the losses due to the moving bed are therefore 31 Pa/m.<br />

Using Table 4-8, apply the Wasp method for calculating the pressure losses of the moving<br />

bed. It will be assumed initially that the –0.044 mm particles are part of the homogeneous<br />

liquid layer above the bed.<br />

It is essential first to determine the drag coefficient and the particle Reynolds number.<br />

2 0.25<br />

����<br />

{log10 [(�/Di)/3.7 + 5.74/Re<br />

) 1244<br />

��<br />

(2) 0.159<br />

0.9 ]} 2<br />

TABLE 4-8 Particle Size versus Volume Concentration in the Slurry (Example 4-11)<br />

New volumetric Volumetric concentration<br />

Original volumetric concentration C V in the <strong>slurry</strong><br />

concentration C V in the solids (at overall solids C V<br />

Particle size (mm) in the solids (after screening) of mixture at 8%)<br />

+0.850 14.3% — —<br />

–0.850 to + 0.400 1.61% 1.88% 0.15%<br />

–0.400 to + 0.200 1.91% 2.23% 0.178%<br />

–0.200 to +0.105 1.41% 1.65% 0.132%<br />

–1.05 to +0.044 1% 1.17% 0.093%<br />

–0.044 79.8% 93.1% 7.45%


4.36 CHAPTER FOUR<br />

Two cases will be considered: spheres and particles with an Albertson sha<strong>pe</strong> factor of 1.0<br />

for the sake of simplicity.<br />

To calculate the particle Reynolds number, the density of 1244 kg/m3 , viscosity of 1.3<br />

mPas, and the s<strong>pe</strong>ed of 1.9 m/s of the carrier fluid are used:<br />

Rep = 1,818,154 dp where dp = the average particle size.<br />

To calculate the drag coefficient of a sphere, the Turton equation (Equation 3.8a) is<br />

used. Results are summarized in Table 4-9.<br />

Wasp et al. (1977) recommend using Durand’s equation for each fraction of solids to<br />

determine the increase in pressure losses due to the moving bed:<br />

gDi(�s – �L)/�L 1.5<br />

�Pbed = 82 �PLCvbed���� V 2 �C�D�<br />

After determining the Darcy friction factor at the pi<strong>pe</strong> diameter of 0.159 m and the<br />

s<strong>pe</strong>ed of 1.9 m/s at a liquid loss of 219 Pa/m, the loss due to each fraction becomes<br />

1.5<br />

�Pbed = 17,490 Cvbed� �<br />

Results of calculations are presented in Table 4-10.<br />

The total friction loss is therefore 240 Pa/m + 151.4 = 391.4. By comparison with the<br />

measured 270Pa/m, the calculations for the bed are higher and can be refined by the<br />

method of concentration using Equation 4-30:<br />

log10� � = –1.8<br />

At 391.4 Pa/m, the equivalent fanning factor is<br />

391.4 = 2 ffV 2<br />

1<br />

�<br />

�C�D�<br />

C Vt � �<br />

CA �KxUf �<br />

�<br />

Di<br />

391.4(0.159)<br />

fN = �� = 0.0069<br />

2 2(1.9 )1,244<br />

To calculate Uf, use Equation 2-25 from Chapter 2:<br />

U = Um�( �f N/ �2�)� = 1.9�(0�.0�0�6�/2�)� = 0.1116 m/s<br />

Assuming Kx = 0.4 and � = 1, we can iterate the results.<br />

TABLE 4-9 Drag Coefficient for Particles in Example 4-11, Assuming<br />

Spherical Sha<strong>pe</strong><br />

Drag coefficient<br />

Particle size Average particle Particle Reynolds Drag coefficient for a particle with<br />

distribution (mm) size (mm) number for a sphere sha<strong>pe</strong> factor of 1<br />

–0.850 to + 0.400 0.63 1145 0.395 0.474<br />

–0.400 to + 0.200 0.3 545 0.545 0.572<br />

–0.200 to +0.105 0.15 272 0.706 0.7413<br />

–1.05 to +0.044 0.07 127 1.02 1.07


HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

TABLE 4-10 Calculated Losses for Each Fraction of Solids in the Moving Bed in the<br />

Lab Test (Example 4-11)<br />

To determine the terminal velocity, we turn to Chapter 3, Equation 3-7:<br />

V t 2 =<br />

C D =<br />

4(� S – � L) gd g<br />

��<br />

3�LV t 2<br />

4 (4.074 – 1.244) 9.81 d g<br />

���<br />

3 (1.244) C D<br />

29.76 d<br />

2 g<br />

V t = �<br />

CD<br />

The iterated pressure loss is 349.7 Pa/m, which is still higher than the measured 270<br />

Pa/m. For further iteration, the fanning factor must be recalculated:<br />

349.7 = 2f fV 2<br />

349.7 (0.159)<br />

ff = �� = 0.00616<br />

2 2 (1.9 ) 1244<br />

Uf = 0.106 m/s<br />

With this new iteration we are converging toward 107 + 240 = 347, which is above the<br />

measured 270Pa/m.<br />

Ellis and Round (1963) indicated that Durand’s equation coefficient of 82 was too<br />

high for nickel sus<strong>pe</strong>nsions. We may therefore divide 270/347 = 0.778 to obtain the new<br />

value of 63.8 for K.<br />

Pi<strong>pe</strong>line Sizing for the Design Flow Rate of 1631 m 3 /hr at a Weight Concentration of 28%<br />

The weight concentration of 28% corresponds to a volumetric concentration of 8.7% and<br />

a mixture density of 1267 kg/m 3 using the solids density of 4074 kg/m 3 . The concentration<br />

of solids in the bed is tabulated in Table 4-11. The flow of 1631 m 3 /hr corresponds to<br />

0.453 m 3 /s.<br />

Consider a 20� OD pi<strong>pe</strong> with a wall thickness of 0.375�, rubber lined with a rubber<br />

thickness of ¼�. The internal diameter of the pi<strong>pe</strong> would be D I = [20 – 2(0.375+0.25)]<br />

= 18.75� or 477 mm. The cross-sectional area of the pi<strong>pe</strong> would be 0.178 m 2 and the average<br />

flow s<strong>pe</strong>ed of the <strong>slurry</strong> would be calculated as V = 0.453/0.178 = 2.55 m/s. Applying<br />

the Thomas–Einstein equation to the volumetric concentration of 8.7% gives an<br />

�<br />

� Di<br />

4.37<br />

Calculated losses Calculated losses for<br />

Particle size Average particle for spherical particles particles with Albertson<br />

distribution (mm) size (mm) (Pa/m) sha<strong>pe</strong> factor of 1.0 (Pa/m)<br />

–0.850 to + 0.400 0.63 58.31 50.87<br />

–0.400 to + 0.200 0.3 53.85 51.93<br />

–0.200 to +0.105 0.15 32.9 31.73<br />

–1.05 to +0.044 0.07 17.56 16.87<br />

Total for bed 162.62 151.4


4.38 CHAPTER FOUR<br />

TABLE 4-11 Iteration for Calculated Losses for Each Fraction of Solids in the<br />

Moving Bed, Based on the Distribution of Concentration—for Lab Tests (Example 4-11)<br />

Average Drag Iterated<br />

particle coefficient Terminal Iterated pressure<br />

Particle size size for a velocity concentration loss<br />

distribution (mm) (mm) sphere (mm/s) –1.8 V t�·K x·U f C/C A (Pa/m)<br />

–0.850 to + 0.400 0.63 0.395 6.89 –0.27 0.537 31.31<br />

–0.400 to + 0.200 0.3 0.545 4.047 –0.163 0.687 36.97<br />

–0.200 to +0.105 0.15 0.706 2.515 –0.1015 0.79 26<br />

–1.05 to +0.044 0.07 1.02 1.43 –0.057 0.877 15.4<br />

Total for bed 109.68<br />

effective viscosity of the mixture of 1.305 mPa · s at 20° C. The pi<strong>pe</strong>line Reynolds number<br />

is therefore<br />

1267(2.55) 0.477<br />

Re = �� = 1,180,931<br />

–3<br />

1.305 × 10<br />

For commercially available rubber-lined pi<strong>pe</strong>s, the roughness is 0.00015 m.<br />

Considering a 477 mm ID pi<strong>pe</strong>, rubber lined, the relative roughness is therefore<br />

0.000315. Applying the Swamee–Jain equation, the Darcy friction factor is calculated as<br />

f D = 0.01578. Loss of carrier fluid is calculated as<br />

0.01578 (2.55 2 ) 1,267<br />

���<br />

2 (0.477)<br />

= 136.3 Pa/m<br />

Using the Wasp method, and applying the Durand’s equation, the calculations yield<br />

�Pbed = 63.8 (136.3)� � 1.5 9.81<br />

��<br />

1 1.5<br />

�Pbed = (18,216) Cvbed�� �C�D� �<br />

The drag coefficient is calculated at the particle Reynolds number using the s<strong>pe</strong>ed of<br />

2.55 m/s, viscosity of 1.305 mPa · s, and density of 1267 kg/m 3 . Re p = 2,475,747 (d p). Results<br />

are presented in Table 4-12.<br />

The Durand equation may then be applied to each fraction of solids. The results are<br />

shown in Table 4-13.<br />

Total losses for <strong>slurry</strong> mixture are therefore calculated as 136.3 + 165.9 = 302 Pa/m.<br />

At 302 Pa/m, the equivalent fanning factor is<br />

302 = 2f f V 2<br />

2.55 2 �C�D�<br />

�<br />

�<br />

Di<br />

302 (0.477)<br />

ff = �� = 0.0089<br />

2 2 (2.55 ) 1244


HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

To calculate Uf, we use Equation 2-15 from Chapter 2:<br />

ff 0.0089<br />

Uf = U�� � = 2.55���<br />

2<br />

2<br />

4.39<br />

Uf = 0.170 m/s<br />

Assuming Kx = 0.4 and � = 1, we can iterate the results based on the distribution of concentration,<br />

as <strong>pe</strong>r Table 4-14. Total friction losses = 136 + 129 = 265 Pa/m or 0.0217<br />

m/m.<br />

TABLE 4-12 Second Iteration for Calculated Losses for Each Fraction of Solids in the<br />

Moving Bed, Based on the Distribution of Concentration—Lab Tests (Example 4-11)<br />

Average Drag Iterated<br />

particle coefficient Terminal Iterated pressure<br />

Particle size size for a velocity concentration loss<br />

distribution (mm) (mm) sphere (mm/s) –1.8 Vt�·K x·Uf C/CA (Pa/m)<br />

–0.850 to +0.400 0.63 0.395 6.89 –0.287 0.516 30.1<br />

–0.400 to +0.200 0.3 0.545 4.047 –0.173 0.671 36.13<br />

–0.200 to +0.105 0.15 0.706 2.515 –0.108 0.78 25.66<br />

–1.05 to +0.044 0.07 1.02 1.43 –0.061 0.868 15.24<br />

Total for bed 107<br />

TABLE 4-13 Drag Coefficient of the Solids in the Pi<strong>pe</strong>line (Example 4-11)<br />

Particle size Average particle Particle Reynolds Drag coefficient<br />

Drag coefficient<br />

for a particle with<br />

distribution (mm) size (mm) number for a sphere sha<strong>pe</strong> factor of 1<br />

–0.850 to + 0.400 0.63 1547 0.414 0.497<br />

–0.400 to + 0.200 0.3 743 0.493 0.52<br />

–0.200 to +0.105 0.15 384 0.602 0.632<br />

–1.05 to +0.044 0.07 186 0.827 0.861<br />

TABLE 4-14 Calculated Loss for Each Fraction of Solids in the Moving Bed in the<br />

20� Pi<strong>pe</strong>line (Example 4-11)<br />

Volumetric Calculated losses<br />

Drag coefficient concentration in for particles<br />

Average for a particle the <strong>slurry</strong> (at (with the Albertson<br />

Particle size particle size with sha<strong>pe</strong> overall solids C V sha<strong>pe</strong> factor of<br />

distribution (mm) (mm) factor of 1 of mixture at 8.7%) 1.0 (Pa/m)<br />

–0.850 to + 0.400 0.63 0.497 0.164% 50.47<br />

–0.400 to + 0.200 0.3 0.52 0.194% 57.71<br />

–0.200 to +0.105 0.15 0.632 0.144% 37<br />

–1.05 to +0.044 0.07 0.861 0.102% 20.79<br />

Total for bed 165.97


4.40 CHAPTER FOUR<br />

TABLE 4-15 Iteration for Calculated Losses for Each Fraction of Solids in the<br />

Moving Bed, Based on the Distribution of Concentration—for 20� Pi<strong>pe</strong>line<br />

(Example 4-11)<br />

Average Drag Iterated<br />

particle coefficient Terminal Iterated pressure<br />

Particle size size for a velocity concentration loss<br />

distribution (mm) (mm) sphere (mm/s) –1.8 V t�·K x·U f C/C A (Pa/m)<br />

–0.850 to +0.400 0.63 0.497 6.14 –0.163 0.687 34.7<br />

–0.400 to +0.200 0.3 0.52 4.14 –0.1093 0.777 44.85<br />

–0.200 to 0.105 0.15 0.632 2.65 –0.07 0.851 31.5<br />

–1.05 to +0.044 0.07 0.861 2.42 –0.063 0.86 17.88<br />

Total for bed 128.93<br />

The purpose of Example 4-11 was to demonstrate the method develo<strong>pe</strong>d by Wasp. A<br />

number of pi<strong>pe</strong>lines have been constructed around the world using this technique and the<br />

practical engineer needs to be familiar with this method as well as with the two-layer<br />

model and stratified flow models that we will explore later.<br />

The following computer program is based on this methodology.<br />

CLS<br />

DIM dp(50), cvdp(50), rep(50), vt(50), cvn(50), dpbed(50), cd(50)<br />

DIM cvind(50), dpav(50), z(50), cca(50), dpnew(50), dfbed(50)<br />

pi = 4 * ATN(1)<br />

DEF fnlog10 (X) = LOG(X) * .4342944<br />

INPUT “name of ore and project”; ore$, proj$<br />

INPUT “date “; dat$<br />

INPUT “your name please “; name$<br />

PRINT “ please choose between the following system of units”<br />

PRINT “ 1- SI units”<br />

PRINT “ 2- US Units”<br />

PRINT<br />

INPUT “ 1 or 2”; ch<br />

10 PRINT<br />

IF rt$ = “Y” OR rt$ = “y” THEN PRINT “<br />


HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

4.41<br />

IF cwe = 1 THEN INPUT “weight concentration in <strong>pe</strong>rcent”; cwin<br />

IF cwe = 2 THEN INPUT “volume concentration in <strong>pe</strong>rcent”; cvin<br />

IF cwe = 0 OR cwe > 2 THEN GOTO 12<br />

PRINT<br />

IF cwe = 1 THEN cw = cwin/100<br />

IF cwe = 1 THEN sgm = sgl/(1 - cw * (1 - sgl/sgs))<br />

IF cwe = 1 THEN cv = (sgm - sgl)/(sgs - sgl)<br />

IF cwe = 1 THEN PRINT USING “s<strong>pe</strong>cific gravity of mixture = ##.##, cv #.###”;<br />

sgm; cv<br />

IF cwe = 2 THEN cv = cvin/100<br />

IF cwe = 2 THEN sgm = cv * (sgs - sgl) + sgl<br />

IF cwe = 2 THEN cw = cv * sgs/sgm<br />

IF cwe = 2 THEN PRINT USING “s<strong>pe</strong>cific gravity of mixture = ##.##, cw =<br />

#.###”; sgm; cw<br />

INPUT “hit any key to continue”; jk$<br />

CLS<br />

PRINT<br />

22 INPUT “pi<strong>pe</strong> outside diameter in inches”; d0<br />

IF d0 = 0 THEN GOTO 22<br />

INPUT “wall thickness in inches”; tw<br />

INPUT “liner thickness in inches”; tl<br />

D1 = d0 - 2 * (tw + tl)<br />

PRINT “inside pi<strong>pe</strong> diameter in inches”; D1<br />

d1m = D1 * .0254<br />

a1 = pi * d1m ^ 2/4<br />

14 v1 = q1m/a1<br />

v1us = v1/.3048<br />

IF rt$ = “Y” OR rt$ = “y” THEN GOTO 18<br />

INPUT “viscosity of <strong>slurry</strong> in cPoise or mPa-s”; viscp<br />

visc = viscp/1000<br />

18 ReL = sgl * 1000 * d1m * v1/visc<br />

PRINT “Reynolds Number of carrier liquid”; ReL<br />

IF rt$ = “Y” OR rt$ = “y” THEN GOTO 20<br />

IF ch = 1 THEN INPUT “pi<strong>pe</strong> roughness in meters”; em<br />

IF ch = 2 THEN INPUT “pi<strong>pe</strong> roughness in feet”; ef<br />

IF ch = 2 THEN em = ef * .3048<br />

edi = em/d1m<br />

PRINT “relative roughness”; edi<br />

20 a = (edi/3.7 + 5.7/ReL ^ .9)<br />

b = fnlog10(a)<br />

fd = .25/b ^ 2<br />

PRINT “darcy factor for carrier liquid”; fd<br />

fan = fd/4<br />

dpl = fd * v1 ^ 2 * sgm * 1000/(2 * d1m)<br />

slopliq = 2 * fan * v1 ^ 2/(9.81 * d1m)<br />

PRINT USING “head loss <strong>pe</strong>r length = ####.##### = “; slopliq<br />

PRINT USING “press drop due to carrier liquid = #####.## Pa/m”; dpl<br />

PRINT<br />

Ub = 9.81 * d1m * (sgm/sgl - 1)/v1 ^ 2<br />

PRINT “Ub”; Ub<br />

INPUT “hit any key to continue”; jk$<br />

INPUT “hit any key to continue”; jk$<br />

PRINT “ starting from the top size you are asked to input particle size for “<br />

PRINT “ each fraction and its volumetric concentration as part of the<br />

solids”<br />

FOR i = 1 TO 50<br />

IF i = 1 THEN PRINT “top fraction”<br />

PRINT “size”; i


4.42 CHAPTER FOUR<br />

IF rt$ = “Y” OR rt$ = “y” THEN GOTO 140<br />

95 INPUT “particle size (in microns) and cumulative volume conc.(%)”; dp(i),<br />

cvdp(i)<br />

140 IF cvdp(i) > 100 THEN GOTO 95<br />

IF i = 1 THEN GOTO 85<br />

cvind(i) = –cvdo + cvdp(i)<br />

dpav(i) = (dpo + dp(i))/2<br />

GOTO 90<br />

85 cvind(i) = cvdp(i)<br />

dpav(i) = dp(i)<br />

90 PRINT USING “average particle size = ###### micron,av.volume conc =<br />

###.#### %”; dpav(i); cvind(i)<br />

cvdo = cvdp(i)<br />

dpo = dp(i)<br />

rep(i) = dpav(i) * 10 ^ -3 * v1 * sgm/visc<br />

PRINT “particle reynolds number”; rep(i)<br />

IF (rep(i) > 2) AND (rep(i) < 500) THEN cd(i) = 18.5 * rep(i) ^ -.6<br />

IF rep(i) < 2 THEN cd(i) = 27 * rep(i) ^ (–.84)<br />

IF (rep(i) > = 500) AND (rep(i) < 200000) THEN cd(i) = .44<br />

vt(i) = (4 * 9.81 * dpav(i) * .000001 * (sgs – sgl)/(3 * cd(i))) ^ .5<br />

IF vt(i) >(v1/2) THEN PRINT “warning deposit velocity is higher than half<br />

the average flow velocity”<br />

PRINT USING “drag coef for av.diam = #.###;terminal velocity = #.#### m/s”;<br />

cd(i); vt(i)<br />

dfbed(i) = 82 * cvind(i) * cv * (Ub ^ 1.5) * (cd(i) ^ -.75)/100<br />

PRINT “ correction to friction factor”; dfbed(i)<br />

‘PRINT “pressure drop due this fraction”; dpbed(i)<br />

hm = hm + dfbed(i)<br />

PRINT USING “ TOTAL correction to friction factor = ###.###”; hm<br />

IF dp(i) = 0 THEN GOTO 130<br />

IF cvdp(i) = 100 THEN GOTO 130<br />

‘INPUT “DO YOU WANT TO CONTINUE (Y/N)”; LKJ$<br />

‘IF LKJ$ = “n” OR LKJ$ = “N” THEN np = i<br />

‘IF LKJ$ = “n” OR LKJ$ = “N” THEN GOTO 130<br />

120 np = i<br />

NEXT i<br />

130 fannew = fan * (1 + hm)<br />

PRINT “total friction factor for <strong>slurry</strong>”; fannew<br />

pressure = fannew * 2 * sgm * 1000 * v1 ^ 2/d1m<br />

slo<strong>pe</strong> = pressure/(9810 * sgm)<br />

PRINT USING “pressure = #####.## Pa/m”; pressure<br />

PRINT “slo<strong>pe</strong> of <strong>slurry</strong> or head <strong>pe</strong>r unit length”; slo<strong>pe</strong><br />

150 INPUT “DO YOU WANT TO DO ITERATION BASED ON CONCENTRATION C/CA (y/n)”; HT$<br />

IF HT$ = “N” OR HT$ = “n” THEN GOTO 325<br />

DFANNEW = fannew<br />

uf = v1 * SQR(DFANNEW/2)<br />

PRINT USING “ FRICTION VELOCITY = ##.### m/s”; uf<br />

FOR i = 1 TO np<br />

IF i = 1 THEN dhm = 0<br />

z(i) = –1.8 * vt(i)/(.38 * uf)<br />

cca(i) = 10 ^ z(i)<br />

PRINT “size,av diam and c/ca “; i, dpav(i), cca(i)<br />

dpnew(i) = dfbed(i) * cca(i)<br />

dhm = dpnew(i) + dhm<br />

NEXT i<br />

DFANNEW = fan * (dhm + 1)<br />

PRINT “REVISED FANNING FACTOR = “; DFANNEW<br />

pressureit = DFANNEW * 2 * sgm * 1000 * v1 ^ 2/d1m


HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

PRINT USING “iterated pressure = #######.## Pa/m = “; pressureit<br />

slo<strong>pe</strong>it = pressureit/(9810 * sgm)<br />

PRINT “revised slo<strong>pe</strong> “; slo<strong>pe</strong>it<br />

325 INPUT “do you want to re<strong>pe</strong>at the calculation for another flow rate<br />

(Y/N)”; rt$<br />

IF rt$ = “Y” OR rt$ = “y” THEN GOTO 10<br />

LPRINT “REVISED FANNING FACTOR = “; DFANNEW<br />

LPRINT USING “iterated pressure = #######.## Pa/m = “; pressureit<br />

LPRINT “revised slo<strong>pe</strong> “; slo<strong>pe</strong>it<br />

525 RETURN<br />

When it was develo<strong>pe</strong>d in the early 1970s, the Wasp method was considered state of<br />

the art. It does, however, ignore or minimize one important parameter, namely the shear<br />

stress between the different su<strong>pe</strong>rimposed layers. This is a topic that the two-layer method<br />

attempts to tackle, as we shall see in Section 4-10. It does, therefore, tend to predict pressure<br />

losses higher than those from stratified flows in certain circumstances of bimodal<br />

(fine and coarse) distribution. Nevertheless, the Wasp method remains a very useful<br />

method to this day for the design of pi<strong>pe</strong>lines, particularly when it is supported by lab<br />

tests, as we showed in Example 4-11.<br />

4-7 SALTATION AND BLOCKAGE<br />

Most modern engineering s<strong>pe</strong>cifications for the design of <strong>slurry</strong> pi<strong>pe</strong>lines categorically<br />

forbid flow at s<strong>pe</strong>eds below V 3. However, the instrumentation engineer needs to know the<br />

pressure rise in saltation or at blockage. Herbich (1991) argued that motors should be<br />

sized to handle the flow in saltation, and there are incidences where it may be economical<br />

to reduce the cross-sectional area of the pi<strong>pe</strong> by allowing flow over a stationary bed.<br />

4.7.1 Pressure Drop Due to Saltation Flows<br />

4.43<br />

In saltation, there is a bed at the bottom part of the horizontal pi<strong>pe</strong> (Figure 4-11) and different<br />

approaches are use to evaluate the pressure losses.<br />

FIGURE 4-11 Concept of Saltation.<br />

Fines in sus<strong>pe</strong>nsion<br />

Saltation bed


4.44 CHAPTER FOUR<br />

Newitt et al. (1955) derived the following equation for saltation flow with d50 > 0.025<br />

mm (0.001 in) and for pi<strong>pe</strong>s smaller than 25.4 mm (1 in):<br />

2 Z = = 66{[�s/�L – 1]gDi/Vm } (4-39)<br />

Babcock (1968) conducted tests on water sus<strong>pe</strong>nsions of coarse sand and steel shot.<br />

He expressed a nondimensional ratio of the square of the s<strong>pe</strong>ed to the product of the pi<strong>pe</strong><br />

inner diameter and the acceleration of gravity. His tests indicated that in the range of 5 <<br />

(V 2 im – i<br />

�<br />

Cvi<br />

/gDi) < 40 the friction losses could be expressed as<br />

�s gDi Z = 60.6� – 1� (4-40)<br />

But for a water sus<strong>pe</strong>nsion of taconite, in the range of 1 < (V2 /gDi) < 12:<br />

�s Z = 66�<br />

gDi – 1� (4-41)<br />

Newitt’s equation (Equation 4-39) is the most commonly used for saltation flows.<br />

Example 4-12<br />

The <strong>slurry</strong> described in Example 4-7 is in saltation at 1.5 m/s. Determine the resultant<br />

friction factor.<br />

Solution in SI Units<br />

Re = = 671,860<br />

0.25<br />

fd1 = ����� = 0.01572<br />

[log10{(0.0002596/3.7) + (5.74/671,860<br />

Using the Newitt equation (4-39), which was derived for small pi<strong>pe</strong>s, would have given<br />

im – iL 1.65 × 9.81 × 0.5778<br />

Z = � = 66����� = 274<br />

2<br />

CviL 1.5<br />

im � – 1 = 60.35<br />

iL<br />

im = 61.35i<br />

Calculating the density of the mixture as:<br />

�m = Cv(�s – �L) + �L = 0.22 (1,650) + 1,000<br />

0.9 )}] 2<br />

1.5 × 1,000 × 0.5778<br />

���<br />

0.00129<br />

�m = 1363 kg/m3 or S.G. = 1.363<br />

Using the Newitt approach with fL = 0.01572<br />

0.01572 × 1.5<br />

im = � � 2<br />

��<br />

�P<br />

� �z<br />

� �L<br />

� �L<br />

2 × 0.5778 × 9.81<br />

� Vm 2<br />

� Vm 2<br />

61.35 = 0.1914 m/m<br />

= � mgi m = 1363 · 9.81 · 0.914 = 2560 Pa/m


HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

A more advanced but less known method to estimate the friction gradient for saltation<br />

flows is the Graf–Acaroglu method that is presented in Chapter 6, which we will also discuss<br />

in the next section with a worked example.<br />

4.7.2 Restarting Pi<strong>pe</strong>lines after Shut-Down or Blockage<br />

4.45<br />

The loss of power, a water hammer situation, or the freezing of a pi<strong>pe</strong>line are occurrences<br />

that require adequate understanding of the power and pressure needed to restart a<br />

pi<strong>pe</strong>line. The concept of the hydraulic radius is defined in Chapter 6 and is essential to<br />

understand blockage.<br />

Vallentine (1955) studied the blockage of a mixture of 0.53 mm (No. 1) and 1.05 mm<br />

(No. 2) sand in 50 mm (2 inch) and 150 mm (6 inch) pi<strong>pe</strong>. He proposed to write the Darcy<br />

equation in terms of a function for blockage:<br />

fD · Q<br />

H = = = �(B) (4-42)<br />

2 fD · Q L<br />

�5 8gD I 2L ��<br />

(8 · g · RH) A2 fD · V 2L ��<br />

(4 · RH) 2 · g<br />

where B is the blocked area of pi<strong>pe</strong> and �(B) is a function of blockage.<br />

Vallentine proposed that the blocked area B is a function of the total flow, pi<strong>pe</strong> diameter,<br />

flow rate of solids, and flow rate of mixture:<br />

1/2 1/3 � L Q m<br />

B = fn� ���<br />

(4-43)<br />

[�s – �L) 1/2 Q<br />

�2.5 1/3<br />

D I Q s<br />

Herbich (1991) plotted these functions. They are presented in Figures 4-12 and 4-13.<br />

In Chapter 6, the work of Graf and Acaroglu is examined in Section 6.5.4. Schiller<br />

(1991) proposed to apply their equation (Equation 6.66b) to the problem of flow over a<br />

BLOCKAGE B<br />

1.0<br />

0.9<br />

0.8<br />

No. 1 Nonuniform (Vallentine)<br />

0.7<br />

No. 2 '' ( “ )<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

Uniform Sands (Craven)<br />

0.0<br />

0 1 2 3 4 5 6 7 8<br />

Q<br />

� d 2.5<br />

�<br />

Q s<br />

��� � –1/3<br />

� �<br />

�s – �w Q<br />

FIGURE 4-12 Blockage factor B. (From Herbich, 1991, reprinted with <strong>pe</strong>rmission from<br />

McGraw-Hill, Inc.)


4.46 CHAPTER FOUR<br />

FIGURE 4-13 Blockage chart. (From Herbich, 1991, reprinted with <strong>pe</strong>rmission from<br />

McGraw-Hill, Inc.)<br />

stationary bed. Schiller established the hydraulic radius of the bed in terms of the ratio of<br />

the mean velocity of flow to the deposition velocity V 3 as<br />

or<br />

BLOCKAGE Q (cfs)<br />

10<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

Q s<br />

V/�(4� ·� g� ·� R� H�)� = V3/�(D� I �· g�)�<br />

R H = 0.25 [(V/V 3) 2 ]D I<br />

Combining these two equations with the Graf–Acaroglu equation (6.64), Schiller proposed<br />

to replace the slo<strong>pe</strong> by head losses <strong>pe</strong>r unit length. Schiller proposed to replace d p<br />

by d 50 for a mixture of particles of different sizes:<br />

{0.25 [(V/V3) 2 ]DI} �[( ���s/��� L�–� 1�)g� d� 3 50�]�<br />

CV · V · = 10.39� � –2.52<br />

�� ���<br />

(4-44)<br />

Example 4-13<br />

Considering that the <strong>slurry</strong> of Example 4.1 is partially blocked at a s<strong>pe</strong>ed of 12 ft/sec. Determine<br />

the resultant head <strong>pe</strong>r unit length to maintain flow:<br />

V3 = 15.25 ft/sec<br />

Solution in SI Units<br />

V = 3.66 m/s<br />

D I = 0.718 m<br />

(�s – �L)d50] �L(h/L) [0.25 [(V/V3) 2 ]DI] The hydraulic radius is therefore [0.25 [(.787) 2 ]0.718] = 0.111 m.<br />

2.0%<br />

� Q 60%<br />

1.0%<br />

0.75%<br />

0.5%<br />

0.25%<br />

0.10%<br />

0.05%<br />

Qs � = Relative Rate of Sediment<br />

Q<br />

Transport<br />

B = Average Pi<strong>pe</strong> Area Blocked<br />

50%<br />

40%<br />

30%<br />

20%<br />

10%<br />

5%<br />

2%<br />

3.00<br />

2.75<br />

2.50<br />

2.25<br />

2.00<br />

1.75<br />

1.50<br />

1.25<br />

1.00<br />

0.75<br />

0.50<br />

PIPE DIAMETER d (ft)


HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

0.27 · 3.66 · 0.111<br />

���<br />

�[(<br />

= 10.39� � –2.52<br />

��<br />

�1�.6�5�)9�.8�1� (�3� � 1�0� (h/L) 0.111<br />

–4 �) 3 �]�<br />

5247 = 10.39[(h/L) 2.52 [839090.5]<br />

h/L = 0.0527<br />

Solution in USCS Units<br />

V = 12 ft/sec<br />

DI = 2.357 ft<br />

The hydraulic radius is therefore [0.25 [(12/15.25) 2 ]2.357] = 0.365 ft.<br />

d50 = 0.000984 ft<br />

0.27 · 12 · 0.365<br />

���<br />

�[(<br />

= 10.39� � –2.52<br />

��<br />

�1�.6�5�)3�2�.2�(0�.0�0�0�9�8�4�) (h/L) 0.365<br />

3 �]�<br />

5256 = 10.39[(h/L) 2.52 [844444]<br />

h/L = 0.0526<br />

Wood 1979 attempted to simplify this complex problem by proposing that the pressure<br />

gradient between the incipient motion velocity V2 (when particles start to leave the<br />

bed) and the actual deposition velocity V3 (when the particles start to move as a heterogeneous<br />

flow) is essentially composed of three components:<br />

1. A component required to overcome the boundary shear stress and turbulence<br />

2. A component required to accelerate the <strong>slurry</strong> due to changes in mean velocity of the<br />

<strong>slurry</strong><br />

3. A component required to accelerate the particles that leave the bed<br />

The third component is the principle source of increase in the pressure gradient. The<br />

particles are assumed to be eroded from the bed at a rate based on flow conditions. If this<br />

rate exceeds the capability of the flow to sus<strong>pe</strong>nd the solids, they tend to fall back into the<br />

bed, until the process is restarted.<br />

The analysis of Wood is based on o<strong>pe</strong>n channel flow, which is the topic of Chapter 6.<br />

Wood treats the stationary bed by considering its hydraulic diameter, and proceeds to apply<br />

a modified Zandi and Govatos approach.<br />

4-8 PSEUDOHOMOGENEOUS<br />

OR SYMMETRIC FLOWS<br />

Above V 4, the flow is pseudohomogeneous or symmetric (but not necessarily uniform).<br />

With negligible hold-up, Govier and Aziz (1972) proposed to express the pressure loss in<br />

terms of the ratio of the friction fanning factor for the <strong>slurry</strong> mixture and the liquid mixture<br />

or as<br />

i m – i<br />

� Cvi L<br />

fNm�m – fNL�L = ��<br />

(4-45)<br />

fNL� L<br />

1.65(3 × 10 –4 )<br />

1.65(0.000984)<br />

4.47


4.48 CHAPTER FOUR<br />

where<br />

fNm = fanning factor of mixtures<br />

fNL = fanning factor of liquid at equivalent volume<br />

In Chapter 1, the increase in dynamic viscosity due to the volumetric concentration of<br />

solids was discussed; it can be expressed as the Einstein–Thomas equation:<br />

� m = � L[1 + AC v + BC v 2 + C exp (DCv)]<br />

where A, B, C, and D are constants. Since the fanning friction factor is a function of<br />

the Reynolds number, the ratio of the Reynolds numbers of the mixture and liquid<br />

would be<br />

Re m =<br />

ReL =<br />

The ratio of the Reynolds number helps to establish a ratio of friction factors:<br />

2 = = {1 + ACv + BCv + C exp(DCv)}� + 1� (4-46)<br />

The next step consists of using the Blasius equation for transition flows:<br />

= {ReL/Rem) 0.25 �LVmDi �<br />

�m<br />

Rem �L Cv(�s – �L) � � ��<br />

ReL �m<br />

�L<br />

fm � (4-47)<br />

fL<br />

Although Govier and Aziz (1972) did not discuss turbulent flow, their analysis can be<br />

extended. For turbulent flows up to Reynolds numbers from 5000 to 100,000,000, which<br />

is well outside the range used for mining, the Swamee–Jain equation (Equation 2.19) can<br />

be used to obtain the ratio of friction factors:<br />

4-9 STRATIFIED FLOWS<br />

� mV mD i<br />

� �m<br />

0.9 2<br />

{log10(0.27 �/Di + 5.74/ReL }<br />

0.9 2<br />

{log10(0.27 �/Di + 5.74/Rem }<br />

fm/fL = ����<br />

(4-49)<br />

In Section 4.6, it was clearly indicated that Wasp et al. (1977) extended the Durand and<br />

Zandi approach to a concept of multiple and su<strong>pe</strong>rimposed layers of particles of different<br />

sizes and volumetric concentration, with a logarithmic concentration distribution. This<br />

method uses the drag coefficient of the particles as an important parameter.<br />

Another school of researchers, particularly lead by Shook, Wilson, and Gilles in Canada,<br />

focused on refining the original models of Newitt, which were based on the terminal<br />

velocity. These authors proposed that the compound mixture of coarse and fine particles<br />

can be simplified to what they called “stratified flows,” in which the fines slide over a<br />

moving bed of coarse solids.<br />

Newitt, as expressed by Equation 4-5, had proposed that the deposition velocity was a<br />

mere factor of 17 times the terminal velocity, Wilson (1991) indicated that this approach<br />

was not confirmed by tests for large pi<strong>pe</strong>s. The concept he proposed was that the flow<br />

was going through a gradual transition from a fully stratified to a fully sus<strong>pe</strong>nded flow,<br />

with a gradual change in the pressure gradient.


Defining a parameter � in terms of the pressure gradient:<br />

� = (4-50)<br />

Moreover, plotting � versus the s<strong>pe</strong>ed, as in Figure 4-14, allows one to find a s<strong>pe</strong>ed<br />

V50 halfway between a fully stratified flow and fully sus<strong>pe</strong>nded flow. The slo<strong>pe</strong> on this<br />

plot is defined as M.<br />

In the region of partially stratified flow, the logarithmic plot confirms that the pressure<br />

gradient � is a function of the velocity raised to the power (–M). M is calculated<br />

from the slo<strong>pe</strong> at V50 considered to be the point at which half-stratified flow occurs, as in<br />

Figure 4-14.<br />

Using V50 as a reference velocity, and considering that the value of � at V50 is effectively<br />

half the mechanical sliding coefficient �p, Wilson (1992) proposed that � = fn(V,<br />

V50, �p, M)<br />

Determining V50 is no easy task. In this partially stratified model, it is argued that the<br />

lifting of particles by turbulence is strongly influenced by their diameter. Thus, the fines<br />

tend to be transported better by the fluid than the coarse solids at the bottom of the pi<strong>pe</strong>.<br />

These particles are transported by eddies, and the largest eddy would be equal to the pi<strong>pe</strong><br />

diameter. The pi<strong>pe</strong> diameter was therefore proposed as a reference. The resistance to motion<br />

of the solid particles is proposed to be the result of a contact load associated with mechanical<br />

sliding friction in the coarser bed and fluid friction associated with the friction<br />

velocity of the carrier fluid. In this complex environment, Wilson et al. (1992) defined a<br />

new settling velocity as<br />

[{�s – �L}g�L] Vs = 0.9 Vt + 2.7� � (4-51)<br />

1/3<br />

im – iL ��<br />

{�m/�L – 1}<br />

��<br />

log (i - i )/( - 1 )<br />

m L m L<br />

HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

p<br />

fully stratified flow<br />

V<br />

50<br />

�L 2/3<br />

partially<br />

stratified<br />

gradient M<br />

S<strong>pe</strong>ed<br />

FIGURE 4-14 Concept of V 50 for stratified flows.<br />

4.49<br />

Fully sus<strong>pe</strong>nded<br />

flow


4.50 CHAPTER FOUR<br />

The two constants 0.9 and 2.7 were obtained from test data on sand. The velocity V 50 is<br />

then expressed in terms of this settling velocity:<br />

V50 = Vs�(8�/f �dL�)� cosh (60 dp/DI) (4-52)<br />

where<br />

cosh = the hy<strong>pe</strong>rbolic trigonometric function<br />

fdL = Darcy friction factor for an equal volume of water<br />

In tests conducted for sand in water at 20° C, � is 0.1 m/s at a particle size of 0.40 mm, indicated<br />

that the magnitude of �p, the mechanical friction coefficient, was 0.44. From Figure<br />

4-14:<br />

� = 0.5 � p(V/V 50) –M = 0.22(V/V 50) –M (4-53)<br />

For a mixture of particles of different diameters, a settling velocity V s85 at d 85 and V s50<br />

at d 50 are used to calculate a standard deviation � s:<br />

Vs85 cosh (60 d85/DI) �s = log����� (4-54)<br />

Vs50 cosh (60 d50/DI) 2 –1/2 The coefficient M is expressed as M = (0.25 + 13�s ) .<br />

The magnitude of Vm/V50 or ratio of mean velocity of the mixture to V50 is defined as the<br />

stratification ratio. The friction losses may be expressed in terms of the stratification ratio.<br />

When d85/d50 < 2, the <strong>slurry</strong> is considered to be narrowly graded, and M is set at 1.7.<br />

For 2 < d85/d50 < 5, 1.7 > M > 0.4.<br />

There is no question that the approach of Wilson et al. (1992) is extremely interesting,<br />

but it is more complex than the methods proposed by Equations 4-3 to 4-7. It requires the<br />

support of testing, a database, a computer software, and a <strong>pe</strong>rsonal computer.<br />

4.10 TWO-LAYER MODELS<br />

Khan and Richardson (1996) explained that Shook and Roco (1991) develo<strong>pe</strong>d a two-layer<br />

model for stratified <strong>slurry</strong> flows, which may be summarized as follows:<br />

� The <strong>slurry</strong> flow of heterogeneous mixtures is considered to consist of two layers, each<br />

with its own velocity of motion and volumetric concentration; but it is assumed that<br />

there is no slip between liquid and solid phases.<br />

� The solids in the up<strong>pe</strong>r layer are fully sus<strong>pe</strong>nded, are at a volumetric concentration<br />

C Vu, and move at a velocityV U.<br />

� The coarser solids in the lower layer are considered to be packed. However, because of<br />

their irregular sha<strong>pe</strong>, there is a certain void fraction between the particles. For sand, the<br />

lower layer is considered to be at a maximum volumetric concentration C VU of 60%<br />

and move at a velocityV B.<br />

The total area of flow for the up<strong>pe</strong>r layer of fines and the lower layer of coarser particles<br />

is A = A B + A U<br />

For the mixture, the mass balance VA = V B A B + V U A U<br />

For the liquid phase, (1 – C V)VA = (1 – C VU) V U A U + (1 – C VB)V B A B<br />

For the solid phase, C VVA = C VBV B A B + C VUV U A U = C VUAV + (C VB – C VU)V B A B


Referring to Figure 4-15:<br />

2 AB = 0.25 D I(� – sin � cos �)<br />

2 AU = 0.25 D I(� – � + sin � cos �)<br />

The up<strong>pe</strong>r wet <strong>pe</strong>rimeter is<br />

WPU = DI(� – �)<br />

WPB = DI� At the interface WPi = DI sin �<br />

The momentum and force balance is expressed as<br />

dP<br />

–AU � = �UWPU + �iWPi for the up<strong>pe</strong>r layer<br />

dx<br />

dP<br />

–AB � = �LWPB – �iWPi + fc�F for the lower layer<br />

dx<br />

dP<br />

–A � = �UWPU + �LWPB + fc�F for the entire pi<strong>pe</strong><br />

dx<br />

fc is the friction factor due to the Coulombic friction between the particles and the pi<strong>pe</strong><br />

and �F is the total force <strong>pe</strong>r unit length exerted normal to the pi<strong>pe</strong> (e.g., weight of the bed<br />

<strong>pe</strong>r unit length, etc.)<br />

2 At low to moderate concentrations �B = 0.5�LL fNBV B. Certain solids are at a concentration<br />

CVB – CVU. They are considered to be of a buoyant weight that is supported at the<br />

wall as a result of interparticle contacts. Averaged over the entire cross-sectional area of<br />

the pi<strong>pe</strong>, these particles define the “contact load” CC, which is calculated as follows:<br />

CC = [CVB – CVU] � (4-55)<br />

A<br />

where<br />

AB = cross-sectional area of the lower layer<br />

A = cross-sectional area of the entire pi<strong>pe</strong><br />

The up<strong>pe</strong>r layer volumetric concentration CVU is obtained from the mean in-situ concentration<br />

CX and the total contact load CC as CVU = Cr – CC. A parameter CX is defined as the in-situ concentration in the x-direction, as<br />

CX = CVB + CVU (4-56)<br />

The relationship between CC and CX is established as an ex<strong>pe</strong>rimental correlation:<br />

where C VB = 0.60.<br />

HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

C C<br />

� CX<br />

= e –� (4-57a)<br />

� = 0.124 Ar –0.061 [(V 2 /(gd p)) 0.28 ][(d p/D i) –0.431 ](� S/� L– 1) –0.272 (4-57b)<br />

And for sand slurries, according to SAC (2000):<br />

� = –0.122 Ar –0.12 [(V/V D) 0.30 ][(d p/D i) –0.51 ](� S/� L– 1) –0.255 (4-57c)<br />

The Archimedean number Ar is defined in Equation 4-5.<br />

A B<br />

4.51


4.52 CHAPTER FOUR<br />

The density of the sus<strong>pe</strong>nded concentration CVU and the density of the carrier liquid �L are then used to compute the shear stresses in the two layers (Figure 4-15).<br />

A fanning friction factor is then obtained for each layer based on some average velocity<br />

of the flow in the pi<strong>pe</strong>, density of the carrier liquid, and dynamic viscosity of the carrier<br />

liquid, and using the relative roughness of the pi<strong>pe</strong> �/Di. The Churchill (1977) equation is then used to calculate the friction factor as<br />

fn = 2[8Re –12 + (A + B) –1.5 ) 1/12 (4-58)<br />

where<br />

A = [–2.457 ln[(7/Re) 0.9 + 0.27 �/Di)] 16<br />

B = (37530/Re) 16<br />

The Reynolds number is calculated on the basis of the average pi<strong>pe</strong> flow s<strong>pe</strong>ed, since<br />

the s<strong>pe</strong>ed of flow in each layer is not known at this point.<br />

An equivalent “sand roughness” is then defined as the ratio of particle size to pi<strong>pe</strong> diameter<br />

(d p/D I). At the interface between the bottom and top layer, a friction factor is derived<br />

by modifying the Colebrook equation (Equation 2-17):<br />

Vertical distance (y)<br />

above bottom quadrant<br />

D I<br />

WP<br />

B<br />

V B V U<br />

S<strong>pe</strong>ed<br />

A U<br />

2<br />

Vertical distance (y)<br />

above bottom quadrant<br />

A<br />

B<br />

WP<br />

U<br />

WP UB<br />

typical real distribution<br />

CVU<br />

Solids volumetric concentration<br />

FIGURE 4-15 Two-layer modeling of coarse and fine particle mixtures.<br />

C VB


HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

fNi = (4-59)<br />

where<br />

Y = 0 for dp/Di < 0.0015<br />

Y = 4 + 1.42 log10 (dp/Di) for 0.0015 < dp/Di < 0.15<br />

and Ar (the Archimedean number) < 3 × 105 2(1 + Y)<br />

���<br />

[4 log10(dp/Di) + 3.36] 2<br />

For bimodal mixtures (mixtures of fine and coarse particles), the Saskatchewan Research<br />

Council (SRC) (2000) suggested that the effects of the fines on the viscosity of the<br />

carrier liquid should be used to calculate the apparent viscosity �f. However, the actual<br />

density of the liquid should be used without corrections for volumetric concentration of<br />

fines.<br />

Khan and Richardson (1996) pointed out that the concept of an equivalent sand roughness<br />

is very s<strong>pe</strong>culative as the transition between one layer and the other. By subsequent<br />

iterations, the friction factor and velocity for each of the two layers are obtained from the<br />

shear stress as the product of the friction factor and the dynamic pressure:<br />

2<br />

�U = 0.5fNU�UVU � B = 0.5f NB� BV B 2<br />

2 2 fNB�BV BWPB D i gfc(�s – �L)(CVL – CVU)(1 – CVB)(sin � – � cos �)<br />

�BWPB = �� + ������ (4-60)<br />

2<br />

2(1 + CVU – CVB) where fc is the coefficient of kinematic friction between the particles and the pi<strong>pe</strong>.<br />

The total normal force <strong>pe</strong>r unit length was defined by Shook as<br />

2 D i g(�s – �L)(sin � – � cos �) (CVB – CVU) (1 – CVB) �F = ��������<br />

(4-61)<br />

2<br />

1 – (CVB – CVU) where � is the half angle formed by the bottom layer with res<strong>pe</strong>ct to the center of the pi<strong>pe</strong>.<br />

To appreciate the complexity of this approach, SRC (2000) indicated that this approach<br />

yielded six unknowns (VB, VU, �, Cr, Cc, and –dP/dx). The numbers of iterations<br />

that are required are better dealt with on a computer.<br />

The two-layer models have gained wide acceptance in the oil–sand industry. The following<br />

example is an illustration at the first level of iteration.<br />

Example 4-14<br />

Sand <strong>slurry</strong> in a pi<strong>pe</strong> is flowing at 6.5 m/s (21.3 ft/sec). The pi<strong>pe</strong> diameter is 717 mm<br />

(28.35�) pi<strong>pe</strong> and the sand particle diameter dp = 360 �m (0.0142�).The volumetric concentration<br />

was presented to be 0.27. Upon review of the composition of the sand, it was<br />

noticed that 15% of the solids were fines smaller than 74 �m. If the lower bed is packed<br />

at 60%, the contact load Cr = 0.30, and the Columbian friction factor fc is 0.50, determine<br />

the pressure gradient (assume water dynamic viscosity 1 cP, and sand S.G. 2.65).<br />

Volumetric concentration in the up<strong>pe</strong>r layer consists essentially of fines:<br />

CVU = 0.15 · 0.27 = 0.0405<br />

CVB = 0.85 · 0.27 = 0.23<br />

By the Einstein–Thomas equation, the dynamic viscosity of the carrier liquid needs to<br />

be corrected for a concentration of 0.0405 in the up<strong>pe</strong>r layer:<br />

� = � L(1 + (2.5 · 0.0405) + (10.05 · 0.0405 2 ) + 0.00273 exp (16.6 · 0.0405)] = 1.123<br />

� L = 1.123cP<br />

4.53


4.54 CHAPTER FOUR<br />

4 × 9.81 (3.6 × 10<br />

Ar = = 666<br />

–4 ) 3 × 1000 (1650)<br />

����<br />

3 (1.123 × 10 –3 ) 2<br />

� = 0.124 Ar –0.061 [(V 2 /(gd p)) 0.28 ][(d p/D i) –0.431 ](� S/� L – 1) –0.272<br />

= 0.124 (666 –0.061 ) (6.5 2 /(9.81 · 3.6 × 10 –4 ) 0.28 ][0.0005 –0.431 ]1.65 –0.272<br />

= 395.52<br />

= e –395.52 = 0<br />

Density of the liquid and fines in the up<strong>pe</strong>r layer is �U = .0401 · 2650 + (1 – .0401) ·<br />

1000 = 1066 kg/m3 CC �<br />

CX<br />

If it is assumed that, due to the void fractions, the lower layer is full at 60%, and since<br />

the volumetric concentration is 0.23, then the area used is 0.23/0.6 = 0.383.<br />

2 2<br />

area = 2[0.25D I (� – 0.5 sin � cos �)] = 0.383 × 0.25�DI<br />

� � 80 degrees � 0.444 �<br />

A B = 0.25 D I 2 (� – sin � cos �) = 0.25 × 0.717 2 x(1.396 – 0.171) = 0.1574 m 2<br />

A U = 0.25 D I 2 (� – � + sin � cos �) = 0.25 × 0.717 2 x(0.556� + 0.171) = 0.246 m 2<br />

WPU = DI(� – �) = 1.252 m<br />

WPB = DI� = 1 m<br />

The total normal force <strong>pe</strong>r unit length was defined by Shook as<br />

0.717 (0.23 – 0.0405) (1 – 0.23)<br />

�F = ����� 2 9.81 (2650 – 1066)(0.985 – 1.39 × 0.1736)<br />

������<br />

= 632 N/m<br />

Since WP B = 1 m,<br />

2<br />

2 fNB�BV BWPB �BWPB = �� + fc �F<br />

2<br />

fNL 1066V<br />

2 �B(1) = + 0.5 · 632 = 533 fNBV B + 316<br />

dp/DI = 0.00035/0.717 = 0.000488, so Y = 0<br />

2 LL(1)<br />

��<br />

2<br />

f NI = 2/[4 log 10(0.717/0.00035) +3.36] 2 = 0.00725<br />

To determine the friction factors for the up<strong>pe</strong>r and lower layers it is required to determine<br />

the s<strong>pe</strong>ed in each layer. To obtain the difference between the velocity in the up<strong>pe</strong>r<br />

and lower layers, it is essential to obtain the shear stress �i between the two layers, or to<br />

make some assumptions and to proceed with further iterations. In a first iteration, it shall<br />

be assumed that VU = 1.1 VLL. AV = ABVB + AUVU 0.25 · � · 0.7172 · 6.5 = 0.1574 m2 (VL)+ 0.246 m2 (1.1VL) VB = 6.14 m/s<br />

VU = 6.75 m/s<br />

1 – (0.23 – 0.0405)


HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

The following calculations require the reader to review some of the concepts of Chapter<br />

6.<br />

The hydraulic diameter (Equation 6-2) for the up<strong>pe</strong>r layer is 4AU/WPU = 4 ×<br />

0.246/1.252 = 0.786 m.<br />

The Reynolds Number for the up<strong>pe</strong>r layer is ReU = 1066 · 6.75 · 0.786/(1.123 · 10 –3 ) =<br />

5,036,209.<br />

The pi<strong>pe</strong> is rubber coated to a roughness of 0.00015 m, �/Di = 0.0002.<br />

Applying Churchill’s equation to the up<strong>pe</strong>r layer:<br />

A = {–2.457 ln[(7/Re) 0.9 + 0.27 �/Di)} 16 = 1.192 · 1022 B = (37530/Re) 16 = 9.045 · 10 –35<br />

fnu = 2[8Re –12 + (A + B) –1.5 ) 1/12 = 2[3 · 10 –80 + 7.684 · 10 –34 ] 1/12 = 0.0035<br />

2 2 �U = 0.5fNU�UV U = 0.5 · 0.0035 · 1066 · 6.75 = 85 Pa<br />

For the lower layer the hydraulic diameter is 4ABWPB = 0.6296 m.<br />

The Reynolds number for the lower layer is ReU = 1066 · 6.14 · 0.6296/(1.123 · 10 –3 )<br />

= 3,669,531.<br />

The pi<strong>pe</strong> is rubber coated to a roughness of 0.00015 m, �/Di = 0.000238.<br />

Applying Churchill’s equation to the up<strong>pe</strong>r layer:<br />

A = {–2.457 ln[(7/Re) 0.9 + 0.27 �/Di)]} 16 = 1.006 · 1021 B = (37530/Re) 16 = 1.433 · 10 –32<br />

fnu = 2[8Re –12 + (A + B) –1.5 ) 1/12 = 2[1.34 · 10 –78 + 3.134 · 10 –32 ] 1/12 = 0.0047<br />

2 2 �B = 0.5fNB�BVB = 0.5 · 0.0047 · 1066 · 6.14 = 94 Pa<br />

2 2 2 2 At the interface �i, 0.5fNi�U(V U – V B) = 0.5 · 0.0035 · 1066 · (6.74 – 6.14 ) = 14.68 Pa.<br />

dP<br />

–AU � = �UWPU + �iWPi = 85 · 1.252 + 14.68 · 14.68 · 0.717 · 0.985 = 117 Pa/m<br />

dx<br />

Since A U = 0.246 m 2 , then dP/dx = 475.6 Pa/m.<br />

Equation 4-60 is not applicable at high concentration of fines, as the <strong>slurry</strong> starts to behave<br />

as a non-Newtonian mixture. For particles with d 50 finer than 74 �m, the method<br />

does not give very reliable results.<br />

For flows above a deposited bed (flows with saltation), SRC (2000) proposed to treat<br />

the up<strong>pe</strong>r layer as a noncircular flow. This means that the hydraulic diameter must be determined<br />

from the wetted area. The difference in roughness between these two surfaces<br />

(up<strong>pe</strong>r surface of the bed) and pi<strong>pe</strong> roughness is not well discussed. The hydraulic diameter<br />

is calculated as D HB = 4A B/(WP B).<br />

It is considered that for pi<strong>pe</strong> flows, the deposit velocity is a function of the pi<strong>pe</strong> diameter<br />

raised to the power of 0.4 for noncircular channels and the friction loss gradient is a<br />

function of the ratio V 2 /D H.<br />

By defining V U as the velocity of the up<strong>pe</strong>r layer and V 3 the critical velocity at which<br />

the bed deposits, the following equation is established:<br />

V U<br />

� V3<br />

0.4 DHB 4.55<br />

= � (4-62)<br />

0.4 D


4.56 CHAPTER FOUR<br />

Friction loss <strong>pe</strong>r unit length<br />

i 1<br />

The ratio of friction gradients is<br />

i m<br />

� i3<br />

A U<br />

deposited bed up<strong>pe</strong>r layer<br />

2 V U D<br />

� � 2 V 3 DHB<br />

= = � � 0.2<br />

(4-63)<br />

where i 1 is pressure gradient at V 1.<br />

At deposition, the flow rate in the up<strong>pe</strong>r layer Q U is lower than the flow rate through<br />

the deposited bed Q B, and from the ratio of velocities as <strong>pe</strong>r Equation 4-57:<br />

Q U<br />

� QB<br />

Q U = A UV U<br />

Q 3 = AV 3<br />

D HB<br />

� D<br />

= � � 0.2<br />

Flow rate<br />

FIGURE 4-16 Flow above a deposited bed—two-layer model.<br />

D<br />

�<br />

DHB<br />

A U<br />

� A<br />

WP<br />

U<br />

WP UL<br />

Deposited solids<br />

(4-64)<br />

d


4.11 VERTICAL FLOW OF COARSE<br />

PARTICLES<br />

Newitt et al. (1961) conducted tests for flows of solids in vertical pi<strong>pe</strong>s. For fine solids<br />

they derived the following empirical equation:<br />

= 0.0037� � 1/2 gDi � �� � 2 Di �S (4-65)<br />

In a vertical flow, it would not be possible to develop dunes, a bed, or saltation. There<br />

is no concentration gradient and the flow may be treated as pseudohomogeneous for friction<br />

loss calculations, as discussed in Section 4-4-3.<br />

Since the flow in a vertical pi<strong>pe</strong> is pseudohomogeneous, a simple instrument to measure<br />

flow rate of <strong>slurry</strong> is the inverted U column, which consists of 4 elbows and 2 vertical<br />

pi<strong>pe</strong> spools (Figure 4-17). The first vertical branch must be sufficiently high to eliminate<br />

entrance effects (previously discussed in Chapter 2). Toward the top of the pi<strong>pe</strong>, a pressure<br />

tap measures the static pressure. Pressure loss occurs through the two elbows. Another<br />

pressure tap is added on the downward section of the pi<strong>pe</strong>. The inverted U column is<br />

calibrated on water and the pressure loss is a function of the flow rate as well as the density<br />

of the <strong>slurry</strong>:<br />

where C d is the discharge coefficient, or<br />

�V<br />

�P = K<br />

2<br />

�<br />

2<br />

2 VmDi Q = � Cd 4�<br />

Q = C dD i�(2� ��P�/�� m�)�<br />

The density of the mixture may be calculated from the input data or measured using a<br />

nuclear radiation density gage.<br />

flow<br />

z A<br />

HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

i m – i L<br />

� iLC v<br />

2<br />

1<br />

� V 2<br />

FIGURE 4-17 Inverted U tube piping for measuring flow of a <strong>slurry</strong> mixture.<br />

� dp<br />

� �L<br />

3<br />

4<br />

z B<br />

4.57


4.58 CHAPTER FOUR<br />

In a vertical flow, in addition to the friction losses as discussed for a pseudohomogeneous<br />

flow, there is a hydrostatic pressure gradient, so that the total pressure<br />

drop is:<br />

2 �P = �mg[�z + fmV m �z/(2gDi)]<br />

�P = pressure loss between two points<br />

�z = height difference between two points<br />

For further reading, see the work of Einstein and Graf (1966) who described such a<br />

flow and a concentration meter for water–sand mixtures.<br />

4-12 INCLINED HETEROGENEOUS FLOWS<br />

Between the two ty<strong>pe</strong>s of horizontal and vertical flows that we discussed in this chapter,<br />

there is a range of inclined flows that are important but are not the subject of extensive<br />

studies. In fact, it may be argued that in many plants, most flows are either horizontal or<br />

vertical, with elbows and fittings between them. An understanding of inclined heterogeneous<br />

flows is essential for certain long overland pi<strong>pe</strong>lines, thickener feed <strong>systems</strong>,<br />

pumpbox feed <strong>systems</strong>, and, in some res<strong>pe</strong>cts, to shed light on o<strong>pe</strong>n channel flows. Until<br />

very recently, attention was focused on uniformly graded slurries.<br />

Worster and Denny (1955) indicated that if the pi<strong>pe</strong> is inclined from the horizontal, the<br />

fraction loss is<br />

where<br />

i � = pressure gradient at �<br />

i m = pressure gradient of the mixture in the horizontal pi<strong>pe</strong><br />

i � = i + (i m – i) cos � (4.65)<br />

This equation suggests that the pressure loss is lower in an inclined pi<strong>pe</strong> than in a<br />

horizontal pi<strong>pe</strong>. It also suggests that the pressure gradient is the same upward or downward.<br />

The ex<strong>pe</strong>rimental work of Kao and Hawang (1979) indicates that this is not correct. In<br />

fact, they noticed that the friction losses for upflows increased up to a certain magnitude<br />

of the angle of inclination and then decreased. In the case of downflows, they measured a<br />

reduction of friction losses from the values of the horizontal pi<strong>pe</strong>.<br />

Wilson et al. (1992) have discussed the effect of pi<strong>pe</strong> inclination on their V 50,<br />

and suggest that Worster and Denny’s equation be modified by using (cos �) 1.85 instead<br />

of cos �. They also published data on certain particles with a diameter between 1 mm<br />

and 6 mm. The test data indicated that the Durand factor for deposition velocity F L (see<br />

Equation 4-2) increased up to an angle of inclination of 30 degrees and by as much as<br />

38%. They also noticed that the deposition velocity V 3 increased by 50% at 30 degrees<br />

pi<strong>pe</strong> inclination, but then they noticed a drop at an angle of 40 degrees. They did not<br />

conduct further tests. Interestingly, they noticed a reduction of the Durand factor F L by<br />

0.3 at a negative inclination of 20 degrees. There is a dearth of information on flow in<br />

inclined pi<strong>pe</strong>s, and as overconservative as it may be, Worster and Denny’s equation<br />

continues to be used. This approach should change, particularly when the angle of inclination<br />

is less than 30 degrees or up to –20 degrees.


HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

4.12.1 Critical Slo<strong>pe</strong> of Inclined Pi<strong>pe</strong>s<br />

4.59<br />

Pi<strong>pe</strong>line have to go through dunes and hills. Certain <strong>slurry</strong> pi<strong>pe</strong>lines may follow the topography<br />

(Figure 4-18) but others require pi<strong>pe</strong> bridges to avoid sedimentation of the <strong>slurry</strong><br />

after a shutdown or power failure (Figure 4-19). A question often raised in designing a<br />

pi<strong>pe</strong>line is determining the critical slo<strong>pe</strong> of the pi<strong>pe</strong> to avoid areas of blockage once the<br />

flow is shut down. To avoid blockage of the line, it is necessary to eliminate areas of steep<br />

slo<strong>pe</strong>s. A commonly used design restriction of 10–16% (5.7–9°) is often adopted when<br />

there is no knowledge of the critical slo<strong>pe</strong>. The idea is that the <strong>slurry</strong> would settle without<br />

segregation to a “soft” consistency and not migrate down to the stee<strong>pe</strong>st slo<strong>pe</strong> in the<br />

pi<strong>pe</strong>line during shutdown.<br />

Kao and Hwang (1979) criticized this approach as being extremely stringent because<br />

it adds to construction costs and capital ex<strong>pe</strong>nses. They implied that it would be better to<br />

pro<strong>pe</strong>rly understand the critical slo<strong>pe</strong> rather than to use a rule of thumb.<br />

Shook et al. (1974) measured the maximum inclination of a 50 mm (2 in) ID Pers<strong>pe</strong>x<br />

pi<strong>pe</strong> on a sand–water mixture. They reported that:<br />

� The maximum rising angle is 14°, or slo<strong>pe</strong> of 24%, before the solids start to slide back.<br />

� The sliding bed was at the interface of the solid bed and the pi<strong>pe</strong> wall rather than within<br />

the settling bed.<br />

FIGURE 4-18 Tailings pi<strong>pe</strong>lines follow the slo<strong>pe</strong> of the hills and use soil friction produced<br />

by partial burial as an anchor.


4.60 CHAPTER FOUR<br />

FIGURE 4-19 Pi<strong>pe</strong>line carrying taconite tailings with an important portion of coarse particles<br />

required pi<strong>pe</strong> bridges to avoid blockage after shutdown or power failure.<br />

� The critical angles for the sand bed increases from 22–26°(slo<strong>pe</strong> of 40–48.7%) with<br />

the decrease of the size of particles. The critical angle is neither affected by the ratio<br />

d p/D I nor by the concentration of the solids.<br />

Durand and Gilbert (1960) derived the following equation for inclined pi<strong>pe</strong>:<br />

i m� = i L + i B(cos �)<br />

where<br />

im = energy gradient for mixture<br />

iL = energy gradient for liquid<br />

iB = energy gradient for solid bed<br />

Kao and Hwang (1979) observed that the critical slo<strong>pe</strong> for glass beads and for sand occurred<br />

at 23°(42% slo<strong>pe</strong>) from the horizontal. For other substances such as coal and walnut<br />

shells, the initial motion ap<strong>pe</strong>ared to occur at the interface between the particle bed<br />

and the pi<strong>pe</strong> wall. This suggested that the internal friction between irregularly sha<strong>pe</strong>d<br />

coarse particles was higher than the friction at the wall of the pi<strong>pe</strong>.<br />

The critical slo<strong>pe</strong> Kao and Hwang (1979) defined was the value for initial particle motion.<br />

For sand or glass beads it was 27° ± 2°, and for coal and walnut shells it was 37° ±<br />

2°.<br />

Craven and Ambrose (1953) investigated the effect of tube inclination on the head loss<br />

for a pi<strong>pe</strong> partially blocked with sediments and for a pi<strong>pe</strong> flowing full. They found that at


HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

4.61<br />

a given average s<strong>pe</strong>ed, an adverse slo<strong>pe</strong> in excess of 10% increased the pressure losses by<br />

25% or more by comparison with a horizontal pi<strong>pe</strong>.<br />

4-12-2 Two-Layer Model for Inclined Flows<br />

Matsouk (1996) develo<strong>pe</strong>d a two-layer model for inclined pi<strong>pe</strong>s. His tests were conducted<br />

in a 150 mm (6�) pi<strong>pe</strong> at angles of inclination between –35 and +35 degrees. The fundamental<br />

equations for this model for the up<strong>pe</strong>r layer are<br />

d(P – �Ugh) �UWPU + �UBWPUB �� = ��<br />

(4-67)<br />

dx<br />

AU<br />

For the lower layer they are<br />

d(P – �Bgh) �BWPB – �iWPi + fc�FN cos � + FW sin �<br />

�� = ����� (4-68)<br />

dx<br />

AL where FW is the submerged weight of the sediments in the lower layer. The force balance<br />

for the whole pi<strong>pe</strong> is then<br />

d(P – �Ugh) �BWPB + �UWPU + fc�FN cos � + FW sin �<br />

�� = ����� (4-69)<br />

dx<br />

A<br />

Equations 4-67 to 4-69 are then solved in a similar manner as presented in Section 4-10.<br />

Matsouk pointed out that his approach was different than the models of Shook and Roco<br />

(1991) (the SRC model), Lazarus (1989), and Lazarus and Cook (1993), which did not include<br />

the pi<strong>pe</strong> axis component of the submerged weight (due to buoyancy) FW sin �. The<br />

tests of Matsouk did not confirm the Worster and Denny equation. At pi<strong>pe</strong> inclinations<br />

close to the angle of internal friction of the transported solids, the behavior of the solids<br />

was different for inclining and descending pi<strong>pe</strong>s under the same s<strong>pe</strong>ed and volumetric<br />

concentration. The difference was larger with coarse than with fine solids. Matsouk con-<br />

z<br />

P<br />

2<br />

L<br />

P + g z<br />

x<br />

P<br />

1<br />

submerged lower layer<br />

of coarse solids<br />

FIGURE 4-20 Concept of the two-layer bipolar flow of <strong>slurry</strong> at an angle of inclination.


4.62 CHAPTER FOUR<br />

cluded that the deformation of the lower layer was essentially due to the submerged<br />

weight of the solids at the bottom of the pi<strong>pe</strong>.<br />

4-13 CONCLUSION<br />

In this chapter, we have shown that two principal schools of heterogeneous <strong>slurry</strong> flows<br />

have develo<strong>pe</strong>d over the last 50 years—one around the Durand–Condolios approach and<br />

the other around the Newitt approach. The former evolved gradually until Wasp modified<br />

it for multilayer compound <strong>systems</strong>. The latter gradually evolved to yield the two-layer<br />

model.<br />

There is no consensus as to which model to use. Some have argued that the Wasp<br />

method was more suitable for coal, whereas the two-layer model was better for sand. This<br />

is based on the number of pa<strong>pe</strong>rs published on two-layer models for sand <strong>slurry</strong> mixtures,<br />

emanating from the great interest in Canadian oil sands.<br />

The science of <strong>slurry</strong> flows is continuously evolving and needs to be understood in the<br />

context of its evolution. It would be a mistake to favor one school of thought over another.<br />

Many successful pi<strong>pe</strong>lines have been designed using either model. The <strong>slurry</strong> engineer<br />

should appreciate that these models are effective tools to be used with correct empirical<br />

coefficients obtained by ex<strong>pe</strong>rimental testing of samples in pumping loops.<br />

After reading this chapter, some readers may get the feeling that designing a <strong>slurry</strong><br />

flow is a combination of science and art. Slurry dynamics may ap<strong>pe</strong>ar to be an exercise of<br />

examining each mixture for its pro<strong>pe</strong>rties, much as the physician must examine each patient<br />

before administering the cure. The flow of coarse particles in water or mixtures of<br />

coarse and fine solids in a liquid is complex. When data is not well accumulated, it is recommended<br />

to conduct <strong>slurry</strong> tests in a pump test loop.<br />

The mixture of coarse and fine particles can lead to a concentration gradient of solids<br />

in horizontal pi<strong>pe</strong>. The coarse particles tend to flow in the lower layers, whereas the fines<br />

flow in the up<strong>pe</strong>r layers. Certain authors recommend determining the pressure losses of<br />

fine solids separately from coarse solids at the corresponding volumetric concentration<br />

and particle diameter of each size. Methods based on concentration ratio for each layer<br />

have been develo<strong>pe</strong>d. A process of iteration is needed to achieve a final estimation of the<br />

pressure loss due to the bed.<br />

New models for inclined flows are ap<strong>pe</strong>aring, such as the work of Matsouk (1996) on<br />

inclined two-layer models The correct design of inclined flow must be based on empirical<br />

data on the critical slo<strong>pe</strong>. The principles reviewed in this chapter apply to dredging and<br />

transporting sand, gravel, coal, steel shot, and rocks from SAG, rod, or ball mills, cyclone<br />

underflows, and tailings, etc., which often have particle sizes larger than 70 �m (mesh<br />

200).<br />

The practical engineer needs to appreciate the limitations of each method and that<br />

models have often been develo<strong>pe</strong>d for a certain range of particle sizes based on ex<strong>pe</strong>rimental<br />

data. It is wise to check the original data.<br />

Because the new methods of stratified flows or two-layer models use the actual hydraulic<br />

diameter of the bed, whereas the Wasp and Durand methods use the actual pi<strong>pe</strong> diameter,<br />

it is easy to get confused. In fact, some of the proponents of the two-layer models<br />

leave the impression that Wasp and Durand are using the “wrong pi<strong>pe</strong> diameter.” This is<br />

not the approach to take. It is wiser to recognize that the Wasp and Durand methods are<br />

useful tools for the range of slurries for which they were develo<strong>pe</strong>d. This includes concentrations<br />

of coarse particles up to a volumetric fraction of 20%. This covers, in fact,<br />

most dredged gravels and sands, coal in a certain range of sizes, as well as crushed rocks.<br />

It is also important to appreciate that the work of Zandi and Govatos (1967) was based on


sands up to a volumetric concentration up to 22%, and it would be erroneous to push the<br />

envelo<strong>pe</strong> of application of their equations beyond such a range.<br />

In the last thirty years, considerable progress has been made with stratified flows. The<br />

new two-layer models push the envelo<strong>pe</strong> of understanding beyond the limits of the models<br />

of Durand, Zandi, and Wasp into the range of volumetric concentrations of 30%. But<br />

these models have their limitations too. Certainly it would be unwise to use the two-layer<br />

model when the d 50 is smaller than 74 micrometers. There is still a considerable amount<br />

of ex<strong>pe</strong>rimental data associated with these stratified models needed to obtain the Conlombic<br />

friction factor and to determine the difference in velocity between the up<strong>pe</strong>r and lower<br />

layers. When the particles are not too coarse, such a difference is not too large, and approximations<br />

such as those proposed by Richardson and Khan are justified, but when the<br />

d 50 is around 400 micrometers or higher, the difference in velocity and shear between the<br />

up<strong>pe</strong>r and lower layers become important.<br />

In Chapter 6, the analysis presented in this chapter will be extended to include o<strong>pe</strong>n<br />

channel flows. The Acaroglu–Graf equation presented in Chapter 6 was applied here to<br />

flows with saltation.<br />

4-14 NOMENCLATURE<br />

a Height of layer A above bottom of conduit<br />

A Cross-sectional area of the entire pi<strong>pe</strong><br />

AB Cross-sectional area of the lower layer<br />

Ap surface area of particle<br />

Ar The Archimedean number<br />

AU Area of up<strong>pe</strong>r layer of flow in the two-layer model<br />

b Factor used to calculate the Archimedean number<br />

B blocked area of pi<strong>pe</strong><br />

C volumetric concentration of the particle diameter under consideration<br />

CA CC CD CE Ct Cv Cvb Concentration of solid particles at a reference plane A (usually at 0.08 DI) Contact load in the Shook–Roco two-layer model<br />

Drag coefficient<br />

Coefficient of discharge<br />

In-situ concentration<br />

Concentration of solids by volume<br />

Volume fraction of solids in the bed<br />

C v bed Concentration of solids in the moving bed<br />

C vi<br />

CVL CVU Cw CX C1 C3 dg dp d85 d50 DH Di dsp HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

4.63<br />

Concentration of solids in the moving bed of fraction i<br />

Concentration of solids by volume in the lower layer in the Shook–Rocco model<br />

Concentration of solids by volume in the up<strong>pe</strong>r layer in the Shook–Rocco model<br />

Weight concentration<br />

In-situ concentration in the two-layer model<br />

Constant<br />

Constant<br />

Diameter of spherical particle<br />

Average diameter of the particle<br />

Sieve passage diameter for 85% of the particles<br />

Sieve passage diameter for 50% of the particles<br />

Hydraulic diameter of a noncircular flow<br />

Inner diameter of pi<strong>pe</strong><br />

diameter of equivalent sphere using the sphericity factor


4.64 CHAPTER FOUR<br />

Es fc fD fDL FL fN fNL fNm fNB fNU Fr<br />

The mass transfer coefficient<br />

coefficient of kinematic friction between particles and pi<strong>pe</strong><br />

Darcy–Weisbach friction factor<br />

Darcy friction factor for equivalent volume of water<br />

Durand–Condolios coefficient to determine the deposition velocity<br />

Fanning friction factor<br />

Fanning friction factor for equivalent volume of water<br />

Fanning friction factor for mixture<br />

Fanning friction factor for the bottom layer in the two-layer model<br />

Fanning friction for the top layer in the two-layer model<br />

Froude number<br />

g Acceleration of falling objects due to gravity (9.78–9.82 m/s2 )<br />

i Equivalent pressure gradient of water at the same volume as the <strong>slurry</strong><br />

im Energy gradient of <strong>slurry</strong> mixture in equivalent m of water <strong>pe</strong>r m of pi<strong>pe</strong> length<br />

i� Pressure gradient for pi<strong>pe</strong> at inclination �<br />

i1 i3 K<br />

Pressure gradient at V1 Pressure gradient at V3 Coefficient or constant<br />

Ke Ex<strong>pe</strong>rimental factor for the pressure gradient<br />

Kf Coefficient in the Durand equation for pressure drop<br />

Kx Von Karman constant<br />

K2 An ex<strong>pe</strong>rimentally determined constant<br />

K3 Coefficient proportional to the mechanical friction factor �<br />

L Length of conduit<br />

Ne Index number<br />

m Power coefficient in Zandi’s models<br />

mi mt M<br />

The mass fraction of solids with particle diameter of dp Total mass of particles<br />

Slo<strong>pe</strong> of the log scale of pressure gradient versus velocity of a stratified flow<br />

P Pressure loss<br />

PW Wetted <strong>pe</strong>rimeter<br />

Q Flow rate<br />

QB Flow rate in the lower layer of the two-layer model<br />

QU Flow rate in the up<strong>pe</strong>r layer of the two-layer model<br />

Ri Inner diameter of a pi<strong>pe</strong><br />

r Local radius for a point in the flow<br />

Re Reynolds number<br />

Rem Reynolds number of the mixture<br />

ReL Reynolds number of the liquid carrier<br />

RH Hydraulic radius<br />

R1 Cross-sectional area of the bed divided by the bed width<br />

s S<strong>pe</strong>cific gravity of the solids<br />

Sf S<strong>pe</strong>cific gravity of liquid<br />

Sm S<strong>pe</strong>cific gravity of <strong>slurry</strong> mixture<br />

U Average velocity<br />

Uf Friction velocity<br />

Ufc Critical friction velocity at which the solids start depositing<br />

Uf 0 Friction velocity at deposition for limiting case of infinite dilution<br />

V Velocity<br />

VB Velocity in the bottom layer<br />

Deposition velocity (also called V3) V D


Vf VL Vm Vmin VR<br />

Velocity of carrier fluid<br />

Velocity of the lower layer in the two-layer model of Shook and Roco<br />

Mean velocity of a mixture<br />

Velocity at minimum pressure drop<br />

Ratio of solid volume concentration of solids to liquid concentration<br />

Vs Settling velocity in the stratified flow model<br />

Vt Terminal velocity of the solid particle<br />

VU Velocity of the up<strong>pe</strong>r layer in the two-layer model of Shook and Roco<br />

V1 Velocity for which flow with a stationary bed is started<br />

V2 Velocity for which flow with a moving bed is started<br />

V3 Deposition velocity or velocity above which solids start to move<br />

V4 Velocity above which all solids move as a pseudohomogeneous mixture<br />

V50 Velocity at 50% stratification<br />

WP Wetted <strong>pe</strong>rimeter<br />

y Distance from the lower boundary (pg 33)<br />

Z Nondimensional parameter to express difference between friction losses due to<br />

the <strong>slurry</strong> and an equivalent volume of water<br />

Subscripts<br />

B Bottom layer<br />

bed Due to the moving bed<br />

i Fraction i<br />

m Mixture<br />

U Up<strong>pe</strong>r layer<br />

Greek letters<br />

� Constant of proportionality<br />

� Roughness<br />

� Angle from the vertical starting at the lowest quadrant point<br />

� The angle from the horizontal<br />

� Pressure loss factor<br />

�r The angle of repose of solid particles<br />

�L Density of liquid carrier<br />

�m Density of the <strong>slurry</strong> mixture<br />

�s Density of solid sediments<br />

�U Density of sus<strong>pe</strong>nded fines and carrier liquid in the up<strong>pe</strong>r layer of the two-layer<br />

model<br />

�W Density of water<br />

�B Shear stress for the lower layer in the two-layer model<br />

�U Shear stress for the up<strong>pe</strong>r layer in the two-layer model<br />

� factor used to compute ratio of concentrations in the two-layer model<br />

� Factor to determine s<strong>pe</strong>ed at 50% stratification in the Wilson model<br />

�L Dynamic viscosity of the carrier liquid<br />

�m Dynamic viscosity of the <strong>slurry</strong> mixture<br />

�p Mechanical friction coefficient<br />

�s Granular stress of the solid particles<br />

� Kinematic viscosity of water<br />

�L Viscosity of carrier liquid at equal volume<br />

� M<br />

HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

Viscosity of <strong>slurry</strong> mixture<br />

� Product of the index number and the volumetric concentration<br />

�s Coefficient of static friction of the solid particles against the wall of the pi<strong>pe</strong><br />

4.65


4.66 CHAPTER FOUR<br />

4-15 REFERENCES<br />

Acaroglu, E. R. 1968. Sediment Transport in Conveyance Systems. Ph.D. dissertation, Cornell University.<br />

Acaroglu, E. R., and W. Graf. 1968. Designing conveyance <strong>systems</strong> for solid–liquid flows. Pa<strong>pe</strong>r<br />

presented at the International Symposium on Solid–Liquid Flow in Pi<strong>pe</strong>s, March 3–7, University<br />

of Pennsylvania, Philadelphia.<br />

Babcock, H. A. 1967. Head losses in pi<strong>pe</strong>line transportation of solids. Pa<strong>pe</strong>r presented at First World<br />

Dredging Conference, WODCON I, the Netherlands.<br />

Babcock, H. A. 1968. Heterogeneous flow of heterogeneous solids. Pa<strong>pe</strong>r presented at International<br />

Symposium on Solid Liquid Flow in Pi<strong>pe</strong>s, March 3–7, University of Pennsylvania, Philadelphia.<br />

Babcock, H. A. 1971. Heterogeneous flow of heterogeneous solids. In I. Zandi (Ed.), Advances in<br />

Solid–Liquid Flows in Pi<strong>pe</strong>s and its Applications. pp. 125–148. Oxford: Pergamon Press.<br />

Bagnold, R. A. 1954. Gravity-free dis<strong>pe</strong>rsion of large spheres in a Newtonian fluid under shear.<br />

Proc. Royal Soc. A, 225, 49–63.<br />

Bagnold, R. A. 1955. Some flume ex<strong>pe</strong>riments on large grains but little denser than the transporting<br />

fluid, and their implications. Proc. Inst. Civ. Eng., 4, 3, 174–205.<br />

Bagnold, R. A. 1957. The flow of cohesionless grains in fluids. Phil. Trans. Roy. Soc., 249, 235–297.<br />

Blatch, N. S. 1906. Water filtration at Washington, DC, discussion trans. Amer. Soc. Civ. Eng. 57,<br />

400–408.<br />

Charles, M. E., and G. S. Stevens. 1972. The pi<strong>pe</strong>line flow of slurries—transitional velocities. Pa<strong>pe</strong>r<br />

presented at the Second International Conference on Hydraulic Transport of Solids and Pi<strong>pe</strong>s.<br />

Second conference of the British Hydromechanic Research Association. Cranfield, England.<br />

Churchill, S. W. 1977. Friction factor equation spans all fluid flow regimes. Chemical Engineering<br />

84, no. 7: 91–92.<br />

Craven, J. P., and H. H. Ambrose. 1953. The transportation of sand in pi<strong>pe</strong>s. Engineering Bulletin<br />

(University of Iowa), 34, 67–88.<br />

Durand R. 1953. Basic Relationship of the transportation of solids in ex<strong>pe</strong>rimental research. Proc. of<br />

the International Association for Hydraulic Research—University of Minnesota, September<br />

1953.<br />

Durand, R., and E. Condolios. 1952. Ex<strong>pe</strong>rimental investigation of the transport of solids in pi<strong>pe</strong>s.<br />

Pa<strong>pe</strong>r presented at Deuxieme Journée de l’hydraulique, Societé Hydrotechnique de France.<br />

Durand, R., and R. Gilbert. 1960. Transport hydraulique et refoulement des mixtures en conduites.<br />

Transactions École des Ponts et Chaussees, 130, 3–4.<br />

Einstein, H. A., and W. H. Graf. 1966. Loop <strong>systems</strong> for measuring sand–water mixtures. Journal of<br />

Hydraulic Division, Am. Soc. Civ. Eng. 92, HY1, pa<strong>pe</strong>r 4608, 1–12.<br />

Ellis, H. S., and G. F. Round. 1963. Laboratory studies on the flow of nickel–water sus<strong>pe</strong>nsions.<br />

Canadian Mining and Metallurgical Bulletin, 56, 773–781.<br />

Faddick R. R. 1982. Ship Loading Coarse Coal Slurries. In The 8th International Conference on Hydraulic<br />

Transport of Solids in Pi<strong>pe</strong>s, Johannesburg, South Africa. Cranfield, UK: BHRA<br />

Group.<br />

Gaessler, H. 1967. Ex<strong>pe</strong>rimentelle und Theoretische Untersuchungen uber die Stromungsvorgange<br />

Beim Transport von Festoffen in Flassigkeiten durch Horizontale Rohrleitungen. Doctoral thesis.<br />

Technische Hochschule, Karlsruhe, Germany. Quoted in G. W. Govier and K. Aziz, The<br />

Flow of Complex Mixtures in Pi<strong>pe</strong>s. New York: Van Nostrand Reinhold Co., 1972, pp. 668<br />

–670.<br />

Geller, L. B, and W. M. Gray. 1986. Selected Theoretical Studies Made in Conjunction with the Joint<br />

Canada/FRG Research Project on Coarse Slurry, Short Distance Pi<strong>pe</strong>line. CANMET SP 86-<br />

16 E–Government of Canada Publications<br />

Gillies, R. G. J. Schaan, R. J. Sumner, M. J. McKibben, and C. A. Shook. 1999. Deposition velocities<br />

for Newtonian slurries in turbulent flows. Pa<strong>pe</strong>r presented at the Engineering Foundation Conference,<br />

Oahu, HI. Submitted for publication in the Canadian J. Chem. Eng. Reference cited by<br />

Saskatchewan Research Council (2000). Slurry pi<strong>pe</strong>line course handout.<br />

Govier, G. W., and K. Aziz. 1972. The Flow of Complex Mixtures in Pi<strong>pe</strong>s. New York: Van Nostrand<br />

Reinhold.


HETEROGENEOUS FLOWS OF SETTLING SLURRIES<br />

4.67<br />

Hayden, J. W., and T. E. Stelson. 1968. Hydraulic conveyance of solids in pi<strong>pe</strong>s. Pa<strong>pe</strong>r presented at<br />

the International Symposium on Solid–Liquid Flow in Pi<strong>pe</strong>s, March 3–7, University of Pennsylvania,<br />

Philadelphia.<br />

Herbich, J. 1991. Handbook of Dredging Engineering. New York: McGraw-Hill.<br />

Hill, R. A., P. E. Snock, and R. L. Gandhi. 1986. Hydraulic transport of solids. In The Pump Handbook,<br />

2nd ed. Edited by I. J. Karassik et al. New York: McGraw-Hill.<br />

Hsu, S. T., A. V. Beken, L. Landweber, and J. F. Kennedy. 1971. The distribution of sus<strong>pe</strong>nded sediment<br />

in turbulent flows in circular pi<strong>pe</strong>s. Pa<strong>pe</strong>r presented at the American Institute of Chemical<br />

Engeering Conference on Solids Transport in Slurries, Atlantic City, NJ.<br />

Hunt, I. N. 1969. Turbulent transport of heterogeneous sediment. Quarterly Journal Mechanics and<br />

App. Maths., 22, 234–246.<br />

Ip<strong>pe</strong>n, A. T. 1971. A new look at sedimentation in turbulent streams. Journal of Boston Soc. Civil<br />

Eng., 58, 3, 131–163.<br />

Ismail, H. M. 1952. Turbulent transfer mechanism and sus<strong>pe</strong>nded sediment in closed channels.<br />

Trans. Amer. Soc. Chem. Eng., 117, 2500, 409–447.<br />

Kao, D. T. Y., and L. Y. Hwang. 1979. Critical slo<strong>pe</strong> for <strong>slurry</strong> pi<strong>pe</strong>lines. Pa<strong>pe</strong>r presented at the Hydrotransport<br />

6 Conference of the British Hydromechanical Research Association, Cranfield,<br />

England.<br />

Khan, A. R., and J. F. Richardson. 1996. Comparison of coarse <strong>slurry</strong> pi<strong>pe</strong>line models. In Proceedings<br />

of Hydrotransport 13, pp. 259–281. Cranfield, UK: BHR Group.<br />

Lazarus, J. H. 1989. Mixed-regime slurries in pi<strong>pe</strong>lines. I. Mechanistic model. Journal of Hydraulic<br />

Engineering, ASCE, 115, 11, 1496–1509.<br />

Lazarus, J. H. & Cooke, R. 1993. Generalised mechanistic model for heterogeneous flow in a non-<br />

Newtonian vehicle. In Proceedings of Hydrotransport 12, pp. 671–690. Cranfield, UK: BHR<br />

Group.<br />

Matsouk V. 1996. Internal structure of <strong>slurry</strong> flow in inclined pi<strong>pe</strong>—Ex<strong>pe</strong>riments and mechanistic<br />

modeling. In Proceedings of Hydrotransport 13, pp. 187–210. Cranfield, UK: BHR Group.<br />

Newitt, D. M., J. F. Richardson, M. Abbott, and R. B. Turtle. 1955. Hydraulic conveying of solids in<br />

horizontal pi<strong>pe</strong>s. Trans Inst. of Chem. Eng., 33, 93–113.<br />

Newitt, D. M., J. R. Richardson, and J. B. Glibbon. 1961. Hydraulic conveying of solids in vertical<br />

pi<strong>pe</strong>s. Trans. Inst. of Chem. Eng., 39, 93–100.<br />

Newitt, D. M., J. R. Richardson, and C. A. Shook (Eds.). 1962. Symposium on Interaction between<br />

Fluids and Particles, London. London: Institution of Chemical Engineers.<br />

Raj, R. S. 1972. Pressure loss in hydraulic transport of solids in inclined pi<strong>pe</strong>s. Pa<strong>pe</strong>r presented at<br />

Hydrotransport 2, Coventry, England.<br />

Saskatchewan Research Council. 2000. Slurry Pi<strong>pe</strong>line Course—SRC Pi<strong>pe</strong> Flow Technology Center,<br />

Saskatoon, Canada, May 15–16.<br />

Schiller, R. E., and P. E. Herbich. 1991. Sediment transport in pi<strong>pe</strong>s. In Handbook of Dredging, Edited<br />

by P. E. Herbich. New York: McGraw-Hill.<br />

Shen, H. W. 1970. Sediment transportation mechanism—Transportation of sediments in pi<strong>pe</strong>s. Journal<br />

Hydraulics Division Am. Soc. Civ. Eng., 96, 1503–1538.<br />

Shook, C. A. 1981. Lead Agency Report for MTCM Coo<strong>pe</strong>rative Research Project. Report E-725-6-<br />

C-81. Report prepared for the Saskatchewan Research Council, Saskatchewan, Canada.<br />

Shook, C. A., J. R. Rollins, and G. S. Vassie. 1974. Sliding in Inclined Slurry Pi<strong>pe</strong>lines and Shutdown.<br />

Report IX. Report prepared for the Saskatchewan Research Council, Saskatchewan,<br />

Canada.<br />

Shook, C. A., and M. C. Roco. 1991. The two layer model. In Slurry Flow: Principles and Practice.<br />

Newton, MA: Butterworth-Heinemann.<br />

S<strong>pe</strong>lls, K. E. 1955. Correlation for use in transport of aqueous sus<strong>pe</strong>nsions of fine solids through<br />

pi<strong>pe</strong>s. Trans Inst. Chem. Eng., 33, 79–84.<br />

Thomas, D. G. 1963. Transport characteristics of sus<strong>pe</strong>nsions: Relation of hindered settling floc<br />

characteristics to rheological parameters. Am. Inst. Chem. Eng. Journal, 9, 310–319.<br />

Thomas, D. G. 1962. Transport Characteristics of sus<strong>pe</strong>nsions, Part IV. Am. Inst. Chem Eng. Journal,<br />

8, 373–378.<br />

Thomas, D. G. 1964. Transport characteristics of sus<strong>pe</strong>nsions, Part IX. Am Inst. Chem. Eng. Journal,<br />

10, 303–308.<br />

Traynis, V. V. 1970.Parameters and Flow Regimes for Hydraulic Transport of Coal by Pi<strong>pe</strong>lines,


4.68 CHAPTER FOUR<br />

Translated and Edited by W. C. Cooley and R. R. Faddick. Terraspace Inc. report, April 1970,<br />

pp 17–19.<br />

Turian, R. M., and T. Yuan. 1977. Flow of slurries and pi<strong>pe</strong>lines. Am. Inst. Chem. Eng. Journal, 23,<br />

3, 232.<br />

Vallentine, H. R.1955. Transportation of Sands in Pi<strong>pe</strong>lines. Commonwealth Engineer (Australia),<br />

April, 349–355.<br />

Warman International Inc. 1990. Slurry Handbook. Madison, WI: Warman International Inc.<br />

Wasp, E. J. et al. 1970. Deposition velocities, transition velocities, and spatial distribution of solids<br />

in <strong>slurry</strong> pi<strong>pe</strong>lines. Pa<strong>pe</strong>r read at 1st International Conference on Hydraulic Transportation of<br />

Solids in Pi<strong>pe</strong>s, Cranfield, England.<br />

Wasp, E. J., J. P. Kenny, and R. L. Gandhi. 1977. Solid–Liquid Flow—Slurry Pi<strong>pe</strong>line Transportation.<br />

Aedermannsdorf, Switzerland: Trans-Tech Publications.<br />

Wicks, M. 1971. Transportation of solids of low concentrations in horizontal pi<strong>pe</strong>s. In Advances in<br />

Solid–Liquid Flow in Pi<strong>pe</strong>s and Its application, edited by I. Zandi. New York: Pergamon Press.<br />

Wilson, K. C. 1970. Slip point of beds in solid–liquid pi<strong>pe</strong> flow. Am. Soc. Chem. Eng. Hydraulic Division,<br />

no. HY1, pa<strong>pe</strong>r 6992: 1–12.<br />

Wilson, K. C. 1991. Pi<strong>pe</strong>line Design for Settling Slurries. In Slurry Handling, edited by N. P. Brown<br />

and N. I. Heywood. New York: Elsevier Applied Sciences.<br />

Wilson, K. C., and D. G. Judge. 1976. Pa<strong>pe</strong>r presented at the International Symposium on Freight<br />

Pi<strong>pe</strong>lines, Washington, DC<br />

Wilson, K. C., G. R. Addie, and R. Clift. 1992. Slurry Transport Using Centrifugal Pumps. New<br />

York: Elsevier Applied Sciences.<br />

Wilson, W. E. 1942. Mechanics of flow of non-colloidal solids. Trans. Am. Soc. of Chem. Eng., 107,<br />

1576.<br />

Wood D. J. 1979. Pressure gradient requirements for re-establishment of <strong>slurry</strong> flow. In Sixth International<br />

Conference on Hydraulic Transport of Solids in Pi<strong>pe</strong>s, p. 217. Cranfield, UK: BHRA<br />

Group.<br />

Worster, R. C., and D. E. Denny. 1955. Hydraulic transport of solid materials in pi<strong>pe</strong>s. Proceedings<br />

of the Institute of Mechanical Engineers (UK), 38, 230–234.<br />

Zandi, I., and G. Govatos. 1967. Heterogeneous flow of solids in pi<strong>pe</strong>line. Proceedings of the Hydraulic<br />

Division of Am. Soc. Civ. Eng., 93, no. HY3, pa<strong>pe</strong>r 5244, 145–159.<br />

Zandi, I. 1971. Hydraulic transport of bulky materials. In Advances in Solid–Liquid Flow in Pi<strong>pe</strong>s<br />

and Its applications, pp. 1–38, I. Zandi (Ed.). Oxford: Pergamon Press.


CHAPTER 3<br />

MECHANICS OF<br />

SUSPENSION OF<br />

SOLIDS IN LIQUIDS<br />

3-0 INTRODUCTION<br />

The physical principles of flow of complex mixtures are based on the interaction between<br />

the different phases, which may mix well or move in su<strong>pe</strong>rimposed layers. In this chapter,<br />

the basic concepts of motion of particles in a carrying fluid will be presented, as well as the<br />

effect of their concentrations and boundaries. In the previous two chapters, we reviewed the<br />

physical pro<strong>pe</strong>rties of solids, single-phase flows, and some as<strong>pe</strong>cts of mixtures of both.<br />

Concepts of non-Newtonian mixtures are reviewed so the reader can understand the<br />

principles used to analyze complex homogeneous flows of very fine particles at high volumetric<br />

concentration.<br />

The physics of solid–liquid mixtures have been the subject of many publications, particularly<br />

by chemical and nuclear engineers. In this chapter, an effort is made to focus on<br />

the practical equations that a <strong>slurry</strong> engineer may use to accomplish his/her tasks. The engineer<br />

may have to use more than one equation when assessing a mixture to make an engineering<br />

judgment.<br />

3-1 DRAG COEFFICIENT AND TERMINAL<br />

VELOCITY OF SUSPENDED SPHERES<br />

IN A FLUID<br />

One fundamental as<strong>pe</strong>ct to the transportation of solids by a liquid is the resistance, called<br />

the drag force, that such solids will exert, and the ability of the liquid to lift such solids,<br />

called the lift force. Both are complex functions of the s<strong>pe</strong>ed of the flow, the sha<strong>pe</strong> of the<br />

solid particles, the degree of turbulence, and the interaction between particles and the<br />

pi<strong>pe</strong>. One approach is to look at a vehicle that we have all come to know—the airplane.<br />

This distraction from the complex world of <strong>slurry</strong> flows is justifiable.<br />

3-1-1 The Airplane Analogy<br />

When an airplane flies in a horizontal plane, it is subject to the forces of downward gravity,<br />

upward lift, and drag opposite to its flight path. To maintain steady flight, its engines<br />

3.1


3.2 CHAPTER THREE<br />

must develop sufficient thrust to overcome drag. The airplane must also fly above its<br />

stalling s<strong>pe</strong>ed.<br />

The lift and drag are aerodynamic forces (Figure 3-1). They are proportional to the<br />

surface area, the density of air, the inclination of the airplane body with res<strong>pe</strong>ct to s<strong>pe</strong>ed,<br />

and the square of the s<strong>pe</strong>ed. For the airplane wing, these forces are expressed as<br />

L = 0.5 CL�V 2Sw (3-1)<br />

D = 0.5 CD�V 2Sw (3-2)<br />

where<br />

� = density of the fluid<br />

V = cruising s<strong>pe</strong>ed of airplane<br />

CL = lift coefficient of wing airfoil<br />

CD = drag coefficient of wing airfoil<br />

The aerodynamic drag consists of two components: the profile drag and induced drag.<br />

The induced drag is proportional to the square of the lift. Airfoils are designed to maximize<br />

the lift-to-drag ratio, or to develop the most lift at the least drag <strong>pe</strong>nalty:<br />

2 CD = CD0 + kwC L (3-3)<br />

where<br />

CD0 = the profile drag<br />

kw = a function of the sha<strong>pe</strong> of the wing (minimum for an elliptical wing and for a wing<br />

flying in ground effect)<br />

The value of the drag and lift coefficients are determined by the sha<strong>pe</strong> of the flying ob-<br />

Thrust<br />

Wing lift<br />

Weight<br />

Forces on an aircraft in<br />

steady horizontal flight<br />

Drag<br />

Stabilizer lift<br />

Weight<br />

Thrust<br />

Drag<br />

Forces on a rocket in<br />

vertical flight<br />

FIGURE 3-1 Lift and drag forces on moving objects.<br />

Buoyancy<br />

Drag<br />

Weight<br />

Forces on a free-falling<br />

particle immersed in a fluid


MECHANICS OF SUSPENSION OF SOLIDS IN LIQUIDS<br />

ject, but also by the physical pro<strong>pe</strong>rties of a fluid, particularly the density, viscosity, and<br />

s<strong>pe</strong>ed of motion. Nondimensional analysis, an important branch of fluid dynamics, allows<br />

the expression of these relationships by characteristic numbers. The Reynolds Number<br />

has already introduced in Chapter 2.<br />

For an airplane in a steady horizontal linear flight, the lift must overcome weight and<br />

the thrust drag. A rocket flying in a vertical plane must develop sufficient thrust to overcome<br />

drag forces as well as weight:<br />

L = W and T = D For an Airplane<br />

T = W + D For a rocket in vertical flight<br />

3-1-2 Buoyancy of Floating Objects<br />

The principle of Archimedes is well known. It states that the buoyancy force develo<strong>pe</strong>d<br />

by an object static in a fluid is equal to the weight of liquid of equivalent volume occupied<br />

by the object. When the density of the object is less than the density of the liquid, the object<br />

floats, and in the inverse situation, the object sinks.<br />

For a sphere immersed in a fluid of density �L, the buoyancy force is calculated from<br />

the weight of fluid the particle displaces:<br />

3 FBF = (�/6)d g�Lg (3-4)<br />

where<br />

FBF = buoyancy force<br />

dg = sphere diameter<br />

g = acceleration due to gravity (9.78–9.81 m/s2 )<br />

3-1-3 Terminal Velocity of Spherical Particles<br />

Although most solids are not spherical in sha<strong>pe</strong>, the sphere is the point of reference for<br />

the analysis of irregularly sha<strong>pe</strong>d solids.<br />

3-1-3-1 Terminal Velocity of a Sphere Falling in a Vertical Tube<br />

When a sphere is allowed to fall freely in a tube, the buoyancy and the drag forces act vertically<br />

upward, whereas the weight force acts downward. At the terminal or free settling<br />

velocity, in the absence of any centrifugal, electrostatic, or magnetic forces<br />

W = D + FBF (3-5)<br />

� �dg 3�Sg = � �dg 3 2<br />

� �<br />

�d<br />

2 g<br />

�Lg + 0.5 CD�LV t� � (3-6)<br />

�<br />

6<br />

�<br />

6<br />

�<br />

4<br />

The drag coefficient corresponding to free fall of the particle is calculated as<br />

4(�S – �L)gdg CD = ��<br />

(3-7)<br />

3�LV t 2<br />

where<br />

d g = sphere diameter<br />

g = acceleration due to gravity, typically 9.8 m/s 2 or 32.2 ft/sec 2<br />

3.3


3.4 CHAPTER THREE<br />

Vt = the terminal (or free settling) s<strong>pe</strong>ed<br />

�s = the density of the solid sphere in kg/m3 or slugs/ft3 �L = the density of the liquid<br />

The terminal (or sinking) velocity is measured using a visual accumulation tube with a<br />

recording drum. Various mathematical models have been derived for the drag coefficient.<br />

Turton and Levenspiel (1986) proposed the following equation:<br />

0.413<br />

0.657 CD = (1 + 0.173Re p ) ���<br />

(3-8)<br />

1 + 1.163 × 104 24<br />

� –1.09<br />

Rep<br />

Re p<br />

Example 3-1<br />

Using the Turton and Levenspiel equation, write a small computer program in quickbasic<br />

to tabulate the drag coefficient of a sphere.<br />

LPRINT “ Drag coefficient vs. Reynolds Number based on<br />

Turton, R., and O. Levenspiel”<br />

RE0= 1<br />

15 FOR I=1 TO 10<br />

RE=I*RE0<br />

CD= (24/RE) * (1+0.173*RE^0.657)*(0.413/(1+11630*RE^-1.09)<br />

PRINT USING “RE= ###### ; Cd = ##.#### “; RE,CD<br />

NEXT I<br />

IF RE>1E6 THEN GOTO 30<br />

RE0=RE<br />

TABLE 3-1 Particle Reynolds Number and Corresponding Drag Coefficient for a<br />

Sphere Based on the Equation of Turton and Levenspiel (1986) as <strong>pe</strong>r Example 3-1<br />

Particle Drag Particle Drag Particle Drag<br />

Reynolds coefficient, Reynolds coefficient, Reynolds coefficient,<br />

number, Rep CD number, Rep CD number, Rep CD 1 28.1520 80 1.2266 6000 0.3983<br />

2 15.2735 90 1.1571 7000 0.4042<br />

3 10.8485 100 1.0994 8,000 0.4151<br />

4 8.5809 200 0.5025 9,000 0.4151<br />

5 7.1908 300 0.6793 10,000 0.4200<br />

6 6.2459 400 0.6085 20,000 0.4497<br />

7 5.5588 500 0.5617 30,000 0.4617<br />

8 5.0349 600 0.5281 40,000 0.4671<br />

9 4.6211 700 0.5029 50,000 0.4697<br />

10 4.2851 800 0.4832 60,000 0.4709<br />

20 2.6866 900 0.4675 70,000 0.4713<br />

30 2.0940 1,000 0.4547 80,000 0.4713<br />

40 1.7729 2,000 0.3990 90,000 0.4711<br />

50 1.5670 3,000 0.3878 100,000 0.4707<br />

60 1.4216 4,000 0.3883 200,000 0.4653<br />

70 1.3124 5,000 0.3927 300,000 0.4609


Drag Coefficient C D<br />

25<br />

20<br />

15<br />

10<br />

5<br />

GOTO 15<br />

30 END<br />

0<br />

0 2 4 6 8 10<br />

Rep<br />

MECHANICS OF SUSPENSION OF SOLIDS IN LIQUIDS<br />

30 0<br />

C D<br />

C D<br />

0<br />

6<br />

4<br />

2<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

20<br />

200<br />

40<br />

400<br />

Results are tabulated in Table 3-1 and presented in Figure 3-2 in a linear scale rather<br />

than a logarithmic scale. Linear scales are sometimes more useful to the mine o<strong>pe</strong>rator<br />

who is in a remote area and has little time to waste on difficult logarithmic graphs<br />

3-1-3-2 Very Fine Spheres<br />

For small particles in the range of a diameter d50 < 0.15 mm (0.0059 in), the most common<br />

equation was created by Stokes and reported by Herbich (1991) and Wasp et al. (1977),<br />

who indicate that the main forces are due to the viscosity effect in the laminar flow regime:<br />

D = 3��dg (3-9)<br />

In the laminar regime, the drag coefficient is inversely proportional to the Reynolds number,<br />

i.e., CD = 24/Rep. The terminal velocity is expressed by Stoke’s equation:<br />

2 (�S – �L)d gg Vt = ��<br />

(3-10)<br />

18�L�<br />

Stokes’s equation is limited to particle Reynolds numbers smaller than 0.1, but has often<br />

been used for particle Reynolds Numbers as large as 1 (based on sphere diameter d g).<br />

60<br />

600<br />

80<br />

800<br />

100<br />

Rep<br />

3<br />

10<br />

Rep<br />

CD CD D C<br />

FIGURE 3-2 Drag coefficient of a sphere for Reynolds number smaller than 300,000.<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0.6<br />

0.4<br />

0.2<br />

1X10 5<br />

2000<br />

4<br />

2X10<br />

4000<br />

4<br />

4X10<br />

Rep<br />

5<br />

3X10<br />

6000<br />

4<br />

6X10<br />

8000<br />

4<br />

8X10<br />

3.5<br />

10 4<br />

Rep<br />

5<br />

1X10<br />

Rep


3.6 CHAPTER THREE<br />

From Equation 3-10, Herbich (1968) pointed out that the radius of particles for which the<br />

validity of the equation is in doubt is expressed as<br />

4.5� 2 � L<br />

� (�S – � L)<br />

R = � � 3/2<br />

This equation is not set in stone for all situations. Rubey (1933) demonstrated one example<br />

by showing that Stoke’s law does not apply to spherical quartz sus<strong>pe</strong>nded in water<br />

when the particle diameter exceeds 0.014 mm (0.00055 in, mesh 105).<br />

3-1-3-3 Intermediate Spheres<br />

For the range of particle Reynolds numbers between 1 and 1000, i.e., when<br />

dpV0 �<br />

1 < � < 1000<br />

�<br />

Govier and Aziz (1972) reported that Allen (1900) derived the following equation:<br />

(� � – � L)g<br />

Vt = 0.2� � 0.72<br />

��<br />

�L<br />

(3-11)<br />

Example 3-2<br />

A <strong>slurry</strong> mixture consists of fine rocks at an average particle diameter of 140 �m, with a<br />

particle density of 2800 kg/m3 . The carrier liquid is water with a dynamic viscosity of 1.5<br />

× 10 –3 Pa · s. The volumetric concentration of the solids is 12%. Determine the terminal<br />

velocity of the particles.<br />

Solution<br />

Using Equation 1-9, the dynamic viscosity of the mixture is<br />

�m = �L[1 + 2.5C� + 10.05C� 2 + 0.00273 exp(16.6C�)]<br />

= 1.5 × 10 –3 [1 + 2.5 × 0.12 + 10.05(0.12) 2 + 0.00273 exp (16.6 × 0.12)]<br />

�m = 2.197 × 10 –3 Pa · s.<br />

Let us check the magnitude of the Reynolds number:<br />

= = 4.468<br />

The Allen law applies in a transition regime:<br />

Vt = 0.2 [9.81 × 1.8] 0.72<br />

(140 × 10 –6 ) 1.18<br />

���<br />

(2.197 × 10 –3 /2800) 0.45<br />

140 × 10 –6 × 0.02504 × 2800<br />

���<br />

2.197 × 10 –3<br />

d�V0� �<br />

�<br />

2.83 × 10<br />

Vt = 0.2 × 7.903<br />

–5<br />

��<br />

0.001789<br />

V t = 0.02504 m/s<br />

d p 1.18<br />

� (�/�) 0.45<br />

Richards (1908) demonstrated that Stokes’s equation is inaccurate for particles with a<br />

diameter larger than 0.2 mm (0.00787 in, mesh 70) and conducted extensive tests for


MECHANICS OF SUSPENSION OF SOLIDS IN LIQUIDS<br />

quartz particles (with a s<strong>pe</strong>cific gravity of 2.65) in laminar, transitional, and turbulent<br />

regimes. He derived the following equation for terminal velocity in mm/s:<br />

8.925<br />

Vt = ����<br />

(3-12)<br />

3 1/2 dg{[1 + 95(�S/�L – 1)d g] – 1}<br />

Where dg, the diameter of the sphere, is expressed in mm. This equation covers the range<br />

of particles between 0.15–1.5 mm (0.0059–0.059 in) at particle Reynolds numbers between<br />

10 and 1000.<br />

3-1-3-4 Large Spheres<br />

For particles with a diameter in excess of 1.5 mm, Herbrich (1991) expressed the terminal<br />

velocity by the following equation:<br />

Vt = Kt�[d� ��� g( ��� S/ L�–� 1�)] � (3-13)<br />

where Kt = an ex<strong>pe</strong>rimental constant = 5.45 for Rep > 800, according to Govier and Aziz<br />

(1972).<br />

Equation 3-13 is often called Newton’s law. In the regime of Newton’s law, the drag<br />

coefficient of a sphere is approximately 0.44, as shown in Figure 3-2. Newton’s law applies<br />

to turbulent flow regimes.<br />

Other equations for terminal velocity of particles have been develo<strong>pe</strong>d by various authors.<br />

Four different equations are presented in Table 3-2.<br />

Example 3-3<br />

Using the Budyruck equation from Table 3-3, determine the terminal velocity of spheres<br />

from 0.1 to1 mm.<br />

A simple computer program is written in quickbasic as follows:<br />

LPRINT<br />

LPRINT “BUDRYCK AND RITTINGER EQUATION FOR TERMINAL<br />

VELOCITY OF SPHERES IN WATER”<br />

LPRINT<br />

LPRINT DP0 = .1<br />

FOR I=1 to 11<br />

DP = I*DP0<br />

VS= (8.925/DP)*(SQR(1+157*DP^30-1)<br />

LPRINT USING “PARTICLE DIAMETER = ##.### mm TERMINAL<br />

VELOCITY Vs = ##.### mm/s”;DP,VS<br />

NEXT I<br />

FOR J=12 TO 20<br />

DP = J*DP0<br />

TABLE 3-2 Equations for Terminal S<strong>pe</strong>ed of Large Spheres<br />

Name Equation* Application<br />

Budryck 3 1/2 Vt = 8.925[(1 + 157d g) – 1]/dg For dg < 1.1 mm<br />

Rittinger Vt = 87(1.65dg) 1/2 For 1.2 < dg < 2 mm<br />

*Where V t is expressed in mm/s and d g in mm.<br />

3.7


3.8 CHAPTER THREE<br />

TABLE 3-3 Calculation of Terminal Velocity of Spheres in Accordance with<br />

Budryck’s Equation<br />

Particle diameter Terminal velocity Particle diameter Terminal velocity<br />

dp in mm Vs in mm/s dp in mm Vs in mm/s<br />

0.1 6.75 0.7 81.63<br />

0.2 22.4 0.8 89.49<br />

0.3 38.34 0.9 96.64<br />

0.4 51.85 1.0 103.26<br />

0.5 63.21 1.1 109.45<br />

0.6 73.02<br />

VS= 87*SQR(1.65*DP)<br />

LPRINT USING “PARTICLE DIAMETER = ##.### mm TERMINAL<br />

VELOCITY Vs = ##.### mm/s”;DP,VS<br />

NEXT J<br />

END<br />

The results are shown in Tables 3-3, 3-4, and Figure 3-3<br />

Herbich (1968) measured drag coefficients for ocean nodules to be as high as 0.6 at<br />

particle Reynolds numbers of 200. This high value is reached with spheres at a particle<br />

Reynolds number of 1000.<br />

3-1-4 Effects of Cylindrical Walls on Terminal Velocity<br />

The previous paragraphs focused on the settling velocity of a single particle or widely<br />

separated particles. The presence of a vessel or cylindrical walls tends to multiply the interaction<br />

between particles and cause some collisions. Extensive tests have been conducted<br />

on flows in vertical tubes. Brown and associates (1950) recommended multiplying the<br />

terminal s<strong>pe</strong>ed of a single particle by a wall correction factor Fw. For laminar flows they<br />

proposed to use the Francis equation:<br />

Fw = 1 – (d�/Di) 9/4 (3-14a)<br />

They proposed to use the Munroe equation for a turbulent flow regime:<br />

Fw = 1 – (d�/Di) 1.5 (3-14b)<br />

where Di = the inner diameter of the tube<br />

TABLE 3-4 Calculation of Terminal Velocity of Spheres in Accordance with<br />

Rittinger’s Equation<br />

Particle diameter Terminal velocity Particle diameter Terminal velocity<br />

d p in mm V t in mm/s d p in mm V t in mm/s<br />

1.1 117.21 1.6 141.36<br />

1.2 122.42 1.7 145.71<br />

1.3 127.42 1.8 149.93<br />

1.4 132.23 1.9 154.04<br />

1.5 136.87 2.0 158.04


in mm/s<br />

Terminal velocity V<br />

t<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

in mm/s<br />

Terminal velocity V<br />

t<br />

MECHANICS OF SUSPENSION OF SOLIDS IN LIQUIDS<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

0 0.01 0.02 0.03 0.04 0.05<br />

Sphere diameter d p in inches<br />

0 0.2 0.4 0.6 0.8 1.0 1.2<br />

Sphere diameter dp<br />

in mm<br />

0.04 0.05 0.06 0.07 0.08<br />

1.0 1.2 1.4 1.6 1.8 2.0<br />

Sphere diameter dp<br />

in mm<br />

(a)<br />

Sphere diameter d p in inches<br />

(b)<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

Terminal velocity V<br />

t<br />

in inch/sec<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

3.9<br />

inch /sec<br />

Terminal velocity V<br />

t<br />

in<br />

FIGURE 3-3 Terminal velocity of spheres (a) in accordance with Budryck’s equation, (b) in<br />

accordance with Rittinger’s equation.


3.10 CHAPTER THREE<br />

Example 3-4<br />

The flow described in Example 3-2 occurs in a 63 mm ID pi<strong>pe</strong>. Determine the corrected<br />

terminal velocity due to the wall effects.<br />

Solution<br />

The terminal velocity was determined to be 0.02504 m/s. The flow is in transition. Equation<br />

3-14a for laminar flow is<br />

Fw = 1 – (d�/DI) 9/4<br />

F w = 1 – (0.140/63) 9/4<br />

Fw = 0.999<br />

Equation 3-14b for turbulent flow is<br />

Fw = 1 – (0.14/63) 1.5 = 0.999.<br />

More recently, Prokunin (1998) extended the analysis of the interaction of the wall<br />

with the motion of a single particle by considering the angle of inclination and any rotation<br />

that the particle may incur. His investigation included immersion in non-Newtonian<br />

flows by testing with glycerin and silicone. He noticed from his tests that when the particle<br />

approaches the wall, it develops a lift force. The lift force seems to increase with a reduction<br />

of the gap that separates the particle from the wall. However, Prokunin could not<br />

explain this lift force and recommended further research.<br />

3-1-5 Effects of the Volumetric Concentration on the<br />

Terminal Velocity<br />

As the volumetric concentration of particles increases, it causes interactions and collisions,<br />

and transfers momentum between the different (finer and coarser) units. The distance<br />

between particles decreases. For spheres at 1% concentration by volume, the interparticle<br />

distance is only 4 diameters. It shrinks to 2.5 diameters at 5% and to 2 diameters<br />

at 10% concentration by volume. In an ideal laminar flow, the interaction is much simpler<br />

than in a turbulent flow.<br />

Worster and Denny (1955) published data on the terminal velocity of coal and gravel<br />

particles, as shown in Table 3-5. The effect of the concentration is clearly marked by a<br />

difference in terminal velocity between a single particle and a volumetric concentration of<br />

30%.<br />

Kearsey and Gill (1963) applied the Carman–Kozeney equation of flow through a<br />

porous medium to determine the terminal velocity as<br />

TABLE 3-5 Terminal Velocity for Coal and Gravel after Worster and Denny (1955)<br />

Coal with a s<strong>pe</strong>cific gravity of 1.5<br />

________________________________<br />

Gravel with a s<strong>pe</strong>cific gravity of 2.67<br />

________________________________<br />

Particle size<br />

____________<br />

Single particle 30% Concentration<br />

______________ ________________<br />

Single particle<br />

______________<br />

30% Concentration<br />

________________<br />

mm Inches (cm/s) (ft/s) (cm/s) (ft/s) (cm/s) (ft/s) (cm/s) (ft/s)<br />

1.59 1/16 4.6 0.15 3.0 0.10 9.1 0.30 3.0 0.10<br />

6.4<br />

12.7<br />

1 –4<br />

1 –2<br />

15.2<br />

30.5<br />

1.50<br />

1.00<br />

10.7<br />

21.3<br />

0.35<br />

0.70<br />

30.5<br />

61.0<br />

1.00<br />

2.00<br />

10.7<br />

21.3<br />

0.35<br />

0.70<br />

25.4 1 51.8 1.70 36.6 1.20 106.7 3.50 36.6 1.20


MECHANICS OF SUSPENSION OF SOLIDS IN LIQUIDS<br />

(1 – Cv) 1 �P<br />

� � 2 �s p L<br />

where<br />

sp = the s<strong>pe</strong>cific surface expressed for as sphere as the surface area to volume ratio:<br />

3<br />

�2 KzC v<br />

V c = � �� �� � (3-15)<br />

�d g 2<br />

3.11<br />

sp = = 6/dg Kz = the Kozney constant, which is a function of particle sha<strong>pe</strong>, porosity, particle orientation,<br />

and size distribution. The magnitude of Kz is between 3 and 6, but is<br />

commonly assumed to be 5<br />

�P/Li = the pressure gradient in the pi<strong>pe</strong> due to the flow of the mixture<br />

In the process of sedimentation, the pressure gradient is essentially due to the volumetric<br />

concentration of the particles and is expressed as<br />

= Cv(�s – �L)g (3-16)<br />

In addition, the settling velocity due to a volumetric concentration is expressed as<br />

Vc = � �� � (3-17)<br />

For spheres with sp = 6/dg, the equation reduces to<br />

Vc = � �� � (3-18)<br />

As the volumetric concentration increases from 3% to 30%, the velocity drops drastically.<br />

Assuming Kz to be equal to 5.0, the settling velocity for spheres reduces to a simple<br />

equation:<br />

(1 – Cv) = (3-19)<br />

where V0 = the terminal velocity at very low volumetric concentration<br />

Equation 3-19 does not apply to volumetric concentrations smaller than 8%. Equation<br />

3-18 would apply to smaller concentrations.<br />

3<br />

(1 – Cv) (�s – �L) �<br />

�<br />

Vc � �<br />

V0 10Cv<br />

3 (1 – Cv) (�s – �L) �2 �s p<br />

2 gd g<br />

��<br />

36KzCv 3 � 3 (�d g/6) �P<br />

�<br />

Li<br />

g<br />

��<br />

KzCv Example 3-5<br />

Assuming that the terminal velocity at a volumetric concentration of 8% is 100 mm/s,<br />

apply Equation 3-18 from a volumetric concentration of 8–30%. Plot the results in Figure<br />

3-4.<br />

Thomas (1963) proposed the following empirical equation in the range of Vc/V0 of<br />

0.08–1.0:<br />

2.303 log10(Vc/V0) = –5.9CV (3-20)<br />

Example 3-6<br />

The free settling s<strong>pe</strong>ed of solid particles is 22 mm/s at a volumetric concentration of 1%.<br />

Using the Thomas equation 3-20, determine the settling s<strong>pe</strong>ed at 25% volumetric concentration.


3.12 CHAPTER THREE<br />

V c/<br />

Vo<br />

Solution<br />

2.303 log 10(V c/V 0) = -5.9 × 0.25<br />

V c/V 0 = 10 –0.64<br />

V c/V 0 = 0.2288<br />

V c = 0.2288 × 22 mm/s = 5.03 mm/s<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 0.1<br />

0.2<br />

Volumetric concentration<br />

FIGURE 3-4 Effect of the volumetric concentration on the terminal velocity of spheres in<br />

accordance with Equation 3-18.<br />

The Kozney-based approach is limited to concentrations where the particles come into<br />

contact with each other in a vertical flow. Beyond this point, the pressure gradient is<br />

smaller than expressed by Equation 3-16. In the case of hard spheres, the settling process<br />

completes when the particles come into contact with each other. In the case of flocculated<br />

particles or clusters of flocculated fluid, stress may cause deformation and further settling<br />

may occur by compaction.<br />

Irregularly sha<strong>pe</strong>d particles and flocculates cause the development of a structure with<br />

its own yield stress level. As the particles move closer, the yield stress increases until<br />

equilibrium is reached. The weight of the overburden is then supported by the saturated<br />

fluid and the compacted sediment.<br />

3-2 GENERALIZED DRAG COEFFICIENT—<br />

THE CONCEPT OF SHAPE FACTOR<br />

Every day the <strong>slurry</strong> engineer has to deal with particles of all sha<strong>pe</strong>s and sizes. Although<br />

the sphere represents a sha<strong>pe</strong> for reference, it is in the minority in the world of crushed or<br />

naturally worn rocks.<br />

Albertson (1953) conducted an extensive study on the effect of the sha<strong>pe</strong> of gravel<br />

particles on the fall velocity in a vertical flow (Figure 3-5). He proposed a definition for a<br />

sha<strong>pe</strong> factor:<br />

where<br />

a = the longest of three mutually <strong>pe</strong>r<strong>pe</strong>ndicular axes<br />

b = the third axis<br />

c = the shortest of three mutually <strong>pe</strong>r<strong>pe</strong>ndicular axes<br />

0.3<br />

c<br />

�A = � (3-21)<br />

�(a�b�)�


MECHANICS OF SUSPENSION OF SOLIDS IN LIQUIDS<br />

b<br />

a<br />

direction of fall<br />

FIGURE 3-5 The axes of an irregularly sha<strong>pe</strong>d particle, according to Albertson.<br />

3.13<br />

Particles in a free fall tend to align themselves to expose the largest surface to the<br />

flow. In other words, they act as free-falling leaves from a tree on an autumn day, where c<br />

is taken as the dimension opposite to the direction of the fall. The projected area of the<br />

particle is a function of the dimensions “a” and “b” but is often not equaled to such a<br />

product as (ab) because particles are usually not rectangular in sha<strong>pe</strong> (see Table 3-6).<br />

In a different approach, Clift et al. (1978) decided to compare the projected area of a<br />

free-falling, irregularly sha<strong>pe</strong>d particle, with a sphere of equal projected area in order to<br />

define a diameter:<br />

da = �(4�S� ���)� f /<br />

(3-22)<br />

where<br />

Sf = the projected area of the free-falling particle<br />

However, Albertson (1953) preferred to define a different diameter base, dp, on the<br />

fact that the actual volume of the free-falling particle could be equated to a sphere of the<br />

TABLE 3-6 Clift Sha<strong>pe</strong> Factor of Various Particles<br />

Isometric Typical mineral particles<br />

____________________________________ _______________________________________<br />

Particle � c Particle � c<br />

Sphere 0.524 Sand 0.26<br />

Cube 0.694 Sillimanite 0.23<br />

Tetrahedron 0.328 Bituminous Coal 0.23<br />

Irregular Rounded 0.54 Blast Furnace Slag 0.19<br />

Cubic angular 0.47 Limestone 0.16<br />

Tetrahedral 0.38 Talc 0.16<br />

Plumbago 0.16<br />

Gypsum 0.13<br />

Flake Graphite 0.023<br />

Mica 0.003<br />

From Wilson et al. (1992).<br />

c


3.14 CHAPTER THREE<br />

same volume but with a diameter of dn. Albertson (1953) therefore proposed a Reynolds<br />

number based on dn: dn�Vt Ren = � (3-23)<br />

�<br />

There may be a marked difference between naturally worn gravel and crushed gravel.<br />

This is a fact that a <strong>slurry</strong> engineer should bear in mind when extrapolating data from lab<br />

results.<br />

Because Clift chose an equivalent diameter d a based on the projected area, he proposed<br />

a different sha<strong>pe</strong> factor:<br />

� c = particle volume/d a 3 (3-24)<br />

Typical values are shown in Table 3-6. The Albertson and Clift sha<strong>pe</strong> factors are about<br />

40 years apart in definition but can be related by a factor E:<br />

� c = E� A<br />

(3-25)<br />

The logarithmic curves as shown in Figure 3-6 are sometimes difficult to read. Table<br />

3-7 presents values of drag coefficient versus Reynolds number rounded off to the first<br />

decimal point.<br />

The work of Albertson was develo<strong>pe</strong>d further by the Inter-Agency Committee on Water<br />

Resources (1958), who develo<strong>pe</strong>d the following two non-dimensional coefficients<br />

(Figure 3-7):<br />

and<br />

Drag coefficient C D<br />

10.0<br />

1.0<br />

0.1<br />

C N = (� s/� L – 1)g�/V t 3 (3-26a)<br />

C N = 0.75C D/Re n<br />

(3-26b)<br />

C S = �(� s/� L – 1)gd p 3 /(6� 2 ) (3-27a)<br />

C S = 0.125�C DRe n 2 (3-27b)<br />

ALBERTSON SHAPE FACTOR = a/ cb<br />

0.3<br />

0.5<br />

0.7<br />

0 10 100 10 3<br />

10 4<br />

10 5<br />

10 6<br />

0 10 100 103 104 105 106 Particle Reynolds number Re p<br />

FIGURE 3-6 The drag coefficient versus Reynolds number and sha<strong>pe</strong> factor. (After Albertson,<br />

1953.)<br />

1.0


MECHANICS OF SUSPENSION OF SOLIDS IN LIQUIDS<br />

TABLE 3-7 Drag Coefficient versus Reynolds Number for Different Albertson Sha<strong>pe</strong><br />

Factors<br />

Drag coefficient<br />

3.15<br />

Reynolds number Sha<strong>pe</strong> factor = 0.3 Sha<strong>pe</strong> factor = 0.5 Sha<strong>pe</strong> factor = 0.7 Sha<strong>pe</strong> factor = 1.0<br />

7 7.0 6.0 4.7 4.0<br />

8 6.5 5.5 4.3 3.7<br />

9 6.1 5.1 4.0 3.4<br />

10 5.8 4.74 3.75 3.15<br />

15 4.64 3.7 3.0 2.4<br />

20 3.95 3.2 2.55 2.0<br />

32 3.0 2.6 2.1 1.55<br />

40 2.7 2.28 1.84 1.3<br />

50 2.5 2.08 1.67 1.12<br />

60 2.3 1.94 1.56 1.0<br />

70 2.25 1.74 1.4 0.94<br />

80 2.2 1.67 1.35 0.844<br />

100 2.08 1.62 1.3 0.8<br />

150 1.87 1.44 1.16 0.68<br />

200 1.75 1.36 1.11 0.6<br />

300 1.74 1.33 1.08 0.5<br />

400 1.8 1.34 1.09 0.44<br />

500 1.9 1.38 1.1 0.4<br />

600 1.94 1.42 1.12 0.38<br />

700 1.988 1.47 1.14 0.36<br />

800 2.0 1.51 1.15 0.34<br />

900 2.07 1.54 1.16 0.334<br />

1000 2.1 1.58 1.17 0.33<br />

2000 2.3 1.72 1.22 0.3<br />

3000 2.28 1.73 1.19 0.29<br />

4000 2.48 1.69 1.16 0.294<br />

5000 2.21 1.66 1.14 0.3<br />

6000 2.2 1.62 1.13 0.31<br />

7000 2.19 1.58 1.13 0.31<br />

8000 2.183 1.55 1.14 0.32<br />

9000 2.18 1.53 1.14 0.32<br />

The drag coefficient C D is then plotted against the equivalent Reynolds number Re n to<br />

determine the terminal velocity. On a logarithmic scale, C N and C S are su<strong>pe</strong>rposed as<br />

straight lines for reference (Figure 3-7).<br />

In order to measure the Albertson sha<strong>pe</strong> factor, Wasp et al. (1977) develo<strong>pe</strong>d a correlation<br />

between the sieve diameter and the fall diameter d n (Figure 3-8).<br />

The approach proposed by Albertson and Clift is limited to free fall of particles in a<br />

fluid. However, turbulence can develop new forces. Whenever an engineering contract requires<br />

the drag of particles to be measured, the engineer is well advised to conduct tests in<br />

a fluid of similar dynamic viscosity as the one that will be used in the project. In addition<br />

to the sha<strong>pe</strong> factor and drag coefficient, the <strong>slurry</strong> engineer must also determine the fluid<br />

density, dynamic viscosity at the tem<strong>pe</strong>rature of pumping, particle density (or s<strong>pe</strong>cific<br />

gravity of solids), nominal (or statistical average) diameter, and fall velocity.


Sieve diameter (mm)<br />

3.16 CHAPTER THREE<br />

FIGURE 3-7 C D and C W versus particle Reynolds number for different sha<strong>pe</strong> factors. Adapted<br />

from the Inter-Agency Committee on Water Resources (1958).<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

spheres<br />

S.F = 0.3<br />

S.F = 0.5<br />

S.F = 0.7<br />

S.F= 0.9<br />

0 0.2 0.4 0.6 0.8 1.0<br />

Sieve diameter (mm)<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

S.F = 0.3<br />

S.F =0.7<br />

S.F =0.5<br />

spheres<br />

0 1 2 3 4<br />

Fall diameter (mm) Fall diameter (mm)<br />

FIGURE 3-8 Relationship between sieve and fall diameter after Wasp et al. (1977).<br />

S.F=0.9<br />

5


MECHANICS OF SUSPENSION OF SOLIDS IN LIQUIDS<br />

Example 3-7<br />

A naturally worn particle has an Albertson sha<strong>pe</strong> factor of 0.7. It has a nominal diameter<br />

of 250 �m. Its density is 3000 kg/m3 . It is allowed to free-fall in water at a tem<strong>pe</strong>rature of<br />

25° C.<br />

Calculate the fall velocity for the single particle and the fall velocity if the volumetric<br />

concentration of particles is increased to 20%.<br />

Solution<br />

Referring to Table 2-7 (or Table 2-8 for USCS units), the kinematic viscosity of water is<br />

0.89 × 10 –6 m2 2 /s. We need to determine the coefficient CS = 0.125�CD/Ren. The curves<br />

published by Inter-Agency Committee on Water Resources indicate that CS =<br />

2 3 2 0.125�CD/Ren = 0.167�(�s/�L – 1)gd p/� = 203.<br />

From Figure 3-6, at a sha<strong>pe</strong> factor of 0.7 and CS of 203, the Reynolds number would<br />

be 7.2Vt = Re/(�dp) = 7.2/(890,000 × 0.00025) = 0.0324 m/s for a single particle.<br />

Applying Equation 3-18 for a concentration of 20%, the velocity would be 0.256 ×<br />

0.0324 = 0.0083 m/s.<br />

3-3 NON-NEWTONIAN SLURRIES<br />

3.17<br />

Various models have been develo<strong>pe</strong>d over the years to classify complex two- and threephase<br />

mixtures (Table 3-8). In the case of mining, the following mixtures are often encountered:<br />

� A fine dis<strong>pe</strong>rsion containing small particles of a solid, which are uniformly distributed<br />

in a continuous fluid and are found in cop<strong>pe</strong>r concentrate pi<strong>pe</strong>lines and in <strong>slurry</strong> from<br />

grinding after classification, etc.<br />

TABLE 3-8 Regimes of Flows for Newtonian and Non-Newtonian Mixtures after<br />

Govier and Aziz (1972)<br />

Single-phase flows<br />

___________________________<br />

Multiphase flows (gas–liquid, liquid–liquid,<br />

gas–solid, liquid–liquid)<br />

___________________________________________________<br />

Single-phase behavior<br />

_____________________________________________________<br />

Multiphase behavior<br />

___________________________<br />

Pseudohomogeneous<br />

_______________________________<br />

Heterogeneous<br />

__________________<br />

True homogeneous Laminar, transition, and<br />

turbulent flow regime<br />

Turbulent flow regime only<br />

Purely viscous Newtonian flows<br />

Purely viscous, non-Newtonian, Bingham plastic<br />

and time-inde<strong>pe</strong>ndent Dilatant<br />

Pseudoplastic<br />

Yield pseudoplastic<br />

Purely viscous, non-Newtonian Thixotropic<br />

and time-de<strong>pe</strong>ndent Rheo<strong>pe</strong>ctic<br />

Viscoelastic Many forms


3.18 CHAPTER THREE<br />

� A coarse dis<strong>pe</strong>rsion containing large particles distributed in a continuous fluid and encountered<br />

in SAG mills, cyclone underflows, and in certain tailings lines, etc.<br />

� A macro-mixed flow pattern containing either a frothy or highly turbulent mixture of<br />

gas and liquid, or two immiscible liquids under conditions in which neither is continuous.<br />

Such patterns are found in flotation circuits in which froth is used to separate concentrate<br />

from gangue.<br />

� A stratified flow pattern containing a gas, liquid, two slurries of different particle sizes,<br />

or two immiscible liquids under conditions in which both phases are continuous.<br />

Designing a pi<strong>pe</strong>line to o<strong>pe</strong>rate in a non-Newtonian flow regime must be based on reliable<br />

test data about the rheology and particle sizing (see Table 3-9). The engineer must<br />

be cautious before venturing into generalizations about rheological pro<strong>pe</strong>rties.<br />

In Figure 1-4 of Chapter 1, the relationship between dynamic viscosity and volumetric<br />

concentration was presented. In fact, the industry has accepted the criterion that friction<br />

losses are highly de<strong>pe</strong>ndent on <strong>slurry</strong> viscosity in cases where the average particle diameter<br />

is finer than 40–60 microns, and (de<strong>pe</strong>nding on the s<strong>pe</strong>cific gravity) at volumetric concentrations<br />

in excess of 30%.<br />

Fibrous slurries such as fermentation broths, fruit pulps, crushed meal animal feed,<br />

tomato puree, sewage sludge, and pa<strong>pe</strong>r pulp may not contain a high <strong>pe</strong>rcentage of solids,<br />

but may flow as non-Newtonian regimes. With these materials, the long fibers are flexible<br />

and intertwine into a close-packed configuration and entrap the sus<strong>pe</strong>nding medium. The<br />

fibers may be flocculated or may form flocs with an o<strong>pe</strong>n structure. Based on the volume<br />

content of the flocs, the mixture may develop high dynamic viscosity. However, because<br />

the flocs are compressible, they may deform with the flow.<br />

Flocculated slurries are encountered in flotation cells circuits, thickeners, and various<br />

processes in mineral extraction plants. With the formation of flocs, the <strong>slurry</strong> may develop<br />

an internal structure. This structure may develop pro<strong>pe</strong>rties leading to a non-Newtonian<br />

flow, shear thinning behavior (pseudoplastic), and sometimes thixotropic time-de<strong>pe</strong>ndent<br />

behavior. When shear stresses are applied to the <strong>slurry</strong>, the floc sizes may shrink and become<br />

less capable of entrapping the carrier <strong>slurry</strong>. At higher shear stresses, the flocs may<br />

shrink to the size of particles, and the flow may lose its non-Newtonian behavior.<br />

3-4 TIME-INDEPENDENT NON-NEWTONIAN<br />

MIXTURES<br />

Certain slurries require a minimum level of stress before they can flow. An example is<br />

fresh concrete that does not flow unless the angle of the chute exceeds a certain minimum.<br />

Such a mixture is said to posses a yield stress magnitude that must be exceeded before<br />

that flow can commence. A number of flows such as Bingham plastics, pseudoplastics,<br />

yield pseudoplastics, and dilatant are classified as time-inde<strong>pe</strong>ndent non-Newtonian fluids.<br />

The relationship of wall shear stress versus shear rate is of the ty<strong>pe</strong> shown in Figure<br />

3-9 (a), and the relationship between the apparent viscosity and the shear rate is shown in<br />

Figure 3-9 (b). The apparent viscosity is defined as<br />

�a = Cw/(d�/dt) (3.28)<br />

3-4-1 Bingham Plastics<br />

For a Bingham plastics it is essential to overcome a yield stress � 0 before the fluid is set in<br />

motion. The shear stress versus shear rate is then expressed as


TABLE 3-9 Examples of Bingham Slurries<br />

MECHANICS OF SUSPENSION OF SOLIDS IN LIQUIDS<br />

Coefficient<br />

Yield of rigidity,<br />

Particle size, Density, Stress, � mPa · s<br />

Slurry d 50 kg/m 3 Pa (cP) Reference<br />

3.19<br />

54.3% Aqueous sus<strong>pe</strong>nsion 92% under 74 �m 1520 3.8 6.86 Hedstrom (1952)<br />

of cement, rock<br />

Flocculated aqueous China 80% under 1 �m 1280 59 13.1 Valentik &<br />

clay sus<strong>pe</strong>nsion No. 1 Whitmore (1965)<br />

Flocculated aqueous China 80% under 1 �m 1207 25 6.7 Valentik &<br />

clay sus<strong>pe</strong>nsion No. 4 Whitmore (1965)<br />

Flocculated aqueous China 80% under 1 �m 1149 7.8 4.0 Valentik &<br />

clay sus<strong>pe</strong>nsion No. 6 Whitmore (1965)<br />

Aqueous clay sus<strong>pe</strong>nsion I 1520 34.5 44.7 Caldwell &<br />

Babitt (1941)<br />

Aqueous clay sus<strong>pe</strong>nsion III 1440 20 32.8 Caldwell &<br />

Babitt (1941)<br />

Aqueous clay sus<strong>pe</strong>nsion V 1360 6.65 19.4 Caldwell &<br />

Babitt (1941)<br />

Fine coal @ 49% C W 50% under 40 �m 1 5 Wells (1991)<br />

Fine coal @ 68% C W 50% under 40 �m 8.3 40 Wells (1991)<br />

Coal tails @ 31% C W 50% under 70 �m 2 60 Wells (1991)<br />

Cop<strong>pe</strong>r concentrate @ 50% under 35 �m 19 18 Wells (1991)<br />

48% C W<br />

21.4% Bauxite < 200�m 1163 8.5 4.1 Boger & Nguyen<br />

(1987)<br />

Gold tails @ 31% C W 50% under 50 �m 5 87 Wells (1991)<br />

18% Iron oxide < 50 �m 1170 0.78 4.5 Cheng &<br />

Whittaker (1972)<br />

7.5 % Kaolin clay Colloidal 1103 7.5 5 Thomas (1981)<br />

Kaolin @ 32% C W 50% under 0.8 �m 20 30 Wells (1991)<br />

Kaolin @ 53% CW with 50% under 0.8 �m 6 15 Wells (1991)<br />

sodium silicate<br />

Kimbelite tails @ 37% C W 50% under 15 �m 11.6 6 Wells (1991)<br />

58% Limestone < 160 �m 1530 2.5 15 Cheng &<br />

Whittaker (1972)<br />

52.4% Fine liminite < 50 �m 2435 30 16 Mun (1988)<br />

Mineral sands tails @ 50% under 160 �m 30 250 Wells (1991)<br />

58% C w<br />

13.9% Milicz clay < 70 �m 2.3 8.7 Parzonka (1964)<br />

16.8% Milicz clay < 70 �m 5.3 13.6 Parzonka (1964)<br />

19.6% Milicz clay < 70 �m 13 25 Parzonka (1964)<br />

Phosphate tails @ 37% C W 85% under 10 �m 28.5 14 Wells (1991)<br />

14% Sewage sludge 1060 3.1 24.5 Caldwell &<br />

Babitt (1941)<br />

Red mud @ 39% C W 5% under 150 �m 23 30 Wells (1991)<br />

Zinc concentrate @ 75% C W 50% under 20 �m 12 31 Wells (1991)<br />

Uranium tails @ 58% C W 50% under 38 �m 4 15 Wells (1991)


3.20 CHAPTER THREE<br />

Shear Stress �<br />

Apparent viscosity � a<br />

Bingham Plastic<br />

Dilatant<br />

Newtonian<br />

Rate of shear (� = du/dy)<br />

Bingham Plastic<br />

Yield Pseudoplastic<br />

Pseudoplastic<br />

Dilatant<br />

Newtonian<br />

Pseudoplastic<br />

Rate of shear (� = du/dy)<br />

(b)<br />

FIGURE 3-9 (a) Shear stress versus shear rate; (b) viscosity versus shear rate of time-inde<strong>pe</strong>ndent<br />

non-Newtonian fluids.


�w – �0 = �d�/dt (3-29)<br />

where<br />

�w = shear stress at the wall<br />

�0 = yield stress<br />

� = the coefficient of rigidity or non-Newtonian viscosity<br />

It is also related to a Bingham plastic limiting viscosity at infinite shear rate by the following<br />

equation:<br />

�0 � = � + �� (3-30)<br />

(d�/dt)<br />

The magnitude of the yield stress �0 may be as low as 0.01 Pascal for sewage sludge<br />

(Dick and Ewing, 1967) or as high as 1000 MPa for asphalts and Bitumen (Pil<strong>pe</strong>l, 1965).<br />

The coefficient of rigidity may be as low as the viscosity of water or as high as 1000 poise<br />

(100 Pa · s) for some paints and much higher for asphalts and bitumen. In the case of tarbased<br />

emulsions or certain tar sands, it is customary to add certain chemicals to reduce the<br />

dynamic viscosity of the emulsion or the coefficient of rigidity of the <strong>slurry</strong>. Tables 3-9<br />

presents examples of Bingham slurries, magnitudes of yield stress, and coefficients of<br />

rigidity � values.<br />

Example 3-8<br />

Samples of a mineral <strong>slurry</strong> with C w = 45% are examined in a lab. From the measurements<br />

of the rate of shear (�) and shear stress (�), determine the yield stress and viscosity.<br />

Rate of Shear � [s –1 ] 100 150 200 300 400 500 600 700 800<br />

Shear Stress � (Pa) 10.93 12.27 13.49 15.68 17.66 19.49 21.2 22.84 24.43<br />

� – � 0 (Pa) 4.11 5.45 6.67 8.87 10.85 12.67 14.39 16.03 17.61<br />

The data is plotted in Figure 3-10. At a low shear rate < 100s – 1, the slo<strong>pe</strong> is<br />

At high shear rate<br />

� = 4.426/100 = 0.0443 Pa · s<br />

4.426<br />

�� = � = 0.0164 Pa · s<br />

270<br />

� =<br />

Take a point at high shear rate (700 s –1 ):<br />

Check at du/dy = 600<br />

at du/dy = 800<br />

MECHANICS OF SUSPENSION OF SOLIDS IN LIQUIDS<br />

� =<br />

� w – � 0<br />

� du/dy<br />

16.03<br />

� 700<br />

� = 0.0229 Pa · s<br />

14.394<br />

� = � = 0.02399<br />

600<br />

3.21


Shear stress (Pa)<br />

3.22 CHAPTER THREE<br />

17.61<br />

� = � 0.022<br />

800<br />

An average � = 0.023 Pa · s is taken.<br />

Alternative � = �0/(du/dy) + �a � = 6.82/700 + 0.0164 = 0.026 Pa · s<br />

This example shows that at zero rate of shear the shear stress is 6.82 Pa. The yield<br />

stress is therefore 6.82 Pa.<br />

The yield stress increases as the concentration of solids augments. Thomas (1961) proposed<br />

the following relationships between yield stress �0, coefficient of rigidity �, concentration<br />

by volume Cv, and viscosity of the sus<strong>pe</strong>nding medium �:<br />

� 0 = K 1C v 3 (3-31)<br />

�/� = exp(K2Cv) (3-32)<br />

where K1 and K2 = constants and are characteristics of the particle size, sha<strong>pe</strong>, and concentration<br />

of the electrolyte concentration.<br />

These equations were derived from the work of Thomas (1961) on sus<strong>pe</strong>nsions of titanium<br />

dioxide, graphite, kaolin, and thorium oxide in a range of particle sizes from<br />

0.35–13 micrometers and in volume concentration of 2–23%.<br />

Thomas (1961) defined a sha<strong>pe</strong> factor �T1 for nonspherical particles as<br />

�T1 = exp[0.7(sp/s0 – 1)] (3-33)<br />

where<br />

sp = the surface area <strong>pe</strong>r unit volume of the actual particles<br />

s0 = the surface area <strong>pe</strong>r unit volume of a sphere of equivalent dimensions or 6/dg He indicated that the coefficient K 1 might then be expressed as<br />

30<br />

28<br />

24<br />

20<br />

16<br />

12<br />

8<br />

4<br />

0<br />

0<br />

0 100 200 300<br />

400 500 600 700<br />

FIGURE 3-10 Plot of data for Example 3-8.<br />

800 900<br />

Rate of shear (sec<br />

-1<br />

)


MECHANICS OF SUSPENSION OF SOLIDS IN LIQUIDS<br />

u�T1 �2 d p<br />

3.23<br />

K 1 = (3-34)<br />

Where K 1 is expressed in Pa (or lb f/ft 2 with u = 210 in), and the particle diameter d p is expressed<br />

in microns.<br />

Thomas defined a second sha<strong>pe</strong> factor � T2 = (s p/s 0) 1/2 to derive the equation:<br />

K 2 = 2.5 + 14� T 2/�d� p� when 0.4 < d p < 20 microns (3-35)<br />

Thomas (1963) extended his work to flocculated mixtures with dis<strong>pe</strong>rsed fine and ultrafine<br />

particles with overall dimensions up to 115 microns. He derived the following equations:<br />

�/� = exp[(2.5 + �)Cv] (3-36)<br />

where<br />

� = �[( �d� f /d� ap� p) � –� 1�]� (3-37)<br />

where<br />

� = the ratio of immobilized dis<strong>pe</strong>rsing fluid to the volume of solids related approximately<br />

to the particle and floc apparent diameter<br />

df = the apparent floc diameter<br />

dapp = the apparent particle diameter<br />

This particle diameter is shown by the following:<br />

dapp = dp(s0/sp) exp(– 1 –<br />

2 ln2 �) (3-38)<br />

where<br />

� = the logarithmic standard deviation<br />

In general, and at a constant tem<strong>pe</strong>rature, the following equations are applied to Bingham<br />

plastic slurries:<br />

�/� = A exp(BCv) (3-39)<br />

�0 = E exp(FCv) (3-40)<br />

The constants A, B, E, and F are derived from tests measuring particle size, sha<strong>pe</strong>, and the<br />

nature of their surface.<br />

Gay et al. (1969) proposed the following correlation for high concentrations of solids:<br />

�/� = exp{[2.5 + [Cv/(Cv� – Cv)] 0.48 ](Cv/Cv�)} (3-41)<br />

where<br />

Cv� = the maximum packing concentration of solids<br />

For a change in tem<strong>pe</strong>rature in the order of 27°C (50°F). Parzonka (1964) develo<strong>pe</strong>d<br />

the following power law equation:<br />

–n � = K3T a (3-42a)<br />

where<br />

n = an exponent<br />

K3 = an exponent<br />

Ta = absolute tem<strong>pe</strong>rature<br />

Govier and Aziz (1972) proposed an equation based on an exponential drop of Bingham<br />

plastic viscosity with tem<strong>pe</strong>rature:


3.24 CHAPTER THREE<br />

� = A exp(B/T) (3-42b)<br />

To obtain the viscosity, plot the curve of the shear stress (� – �0) in Pascals against the<br />

shear rate � (s –1 ).<br />

3-4-2 Pseudoplastic Slurries<br />

Pseudoplastic fluids are time-inde<strong>pe</strong>ndent non-Newtonian fluids that are characterized by<br />

the following:<br />

� An infinitesimal shear stress, which is sufficient to initiate motion<br />

� The rate of increase of shear stress with res<strong>pe</strong>ct to the velocity gradient decreases as<br />

the velocity gradient increases<br />

This ty<strong>pe</strong> of flow is encountered when fine particles form loosely bound aggregates<br />

that are aligned, stable, and reproducible at a given magnitude of shear rate.<br />

The behavior of pseudoplastic fluids is difficult to define accurately. Various empirical<br />

equations have been develo<strong>pe</strong>d over the years and involve at least two empirical factors,<br />

one of which is an exponent. For these reasons, pseudoplastic slurries are often<br />

called power-law slurries. The shear stress is defined in terms of the shear rate by the following<br />

equation:<br />

�w = K[(d�/dt) n ] (3-43)<br />

where<br />

K = the power law consistency factor, expressed in Pa · sn n = the power law behavior index, and is smaller than unity<br />

Examples of pseudoplastic slurries are shown in Table 3-10.<br />

The apparent viscosity of a pseudoplastic is defined in terms of the ratio of the shear<br />

stress to the shear rate:<br />

� a = [� w/(d�/dt)] (3-44)<br />

3-4-2-1 Homogeneous Pseudoplastics<br />

Pseudoplastic slurries are another category of non-Newtonian slurries. Pseudoplastics are<br />

divided into homogeneous and pseudohomogeneous mixtures. Whereas in the case of a<br />

Bingham <strong>slurry</strong>, it was pointed out that the coefficient of rigidity was a linear function of<br />

the shear rate, in the case of a pseudoplastic, the coefficient of rigidity is expressed by the<br />

following power law:<br />

� = K(d�/dt) n–1 (3-45)<br />

The shear stress is plotted against the shear rate on a logarithmic scale at various volume<br />

fractions. From the slo<strong>pe</strong> of such a plot, “K,” the power law consistency factor, and<br />

“n,” the power law behavior index (smaller than unity) are derived as plotted in Figure 3-<br />

11.<br />

As indicated in Figure 3-12 the magnitude “K,” the power law consistency factor, and<br />

the power law factor index n are de<strong>pe</strong>ndent on the volumetric concentration of solids.<br />

Example 3-9<br />

A phosphate <strong>slurry</strong> mixture is tested using a rheogram. The following data describe the<br />

relationship between the wall shear stress and the shear rate:


MECHANICS OF SUSPENSION OF SOLIDS IN LIQUIDS<br />

d�/dt 0 50 100 150 200 300 400 500 600 700 800<br />

�w(Pa) 25 32 43 51 53 56 58 60 62 63.2 64.3<br />

The mixture is non-Newtonian. If it is considered a power law <strong>slurry</strong>, derive the power<br />

law exponent “n” and the power law coefficient K.<br />

Solution<br />

The first step is to plot the data on a logarithmic scale. In the equation for a pseudoplastic,<br />

the coefficient of rigidity is expressed by equations (3.43) and (3.45), the values of “K”<br />

and “n.” By using the logarithmic scale:<br />

log �w = log K + n log (d�/dt)<br />

log(d�/dt) 1.699 2 2.176 2.301 2.477 2.602 2.669 2.778 2.845 2.903<br />

log(�w) 1.505 1.633 1.707 1.724 1.748 1.763 1.778 1.792 1.8 1.808<br />

n — 0.425 0.592 0.136 0.136 0.12 0.154 0.112 0.13 0.14<br />

log(d�/dt) 2 – log(d�/dt) 1<br />

n = ���<br />

(log�w) 2 – (log�w) 1<br />

n � 0.132<br />

1.8 = log K = 0.132 × 2.843<br />

log K = 1.424<br />

K = 26.5<br />

TABLE 3-10 Examples of Power Law Pseudoplastics<br />

Range of Range of Angle of<br />

Particle weight consistency flow<br />

size, concentration, coefficient K, behavior<br />

Slurry d 50 % Ns n /m 2 index, n Reference<br />

3.25<br />

Cellulose acetate 1.5–7.4 1.4–34.0 0.38–0.43 Heywood (1996)<br />

Drilling mud—barite 14.7 �m 1.0–40.0 0.8–1.3 0.43–0.62 Heywood (1996)<br />

Sand in drilling mud 180 �m 1.0–15% 0.72–1.21 0.48–0.57 Heywood (1996)<br />

sand using<br />

drilling mud<br />

with 18%<br />

barite<br />

Graphite 16.1 �m 0.5–5.0 Unknown Probably 1 Heywood (1996)<br />

Graphite and 5 �m 32.2 total<br />

magnesium (4.1 graphite 5.22 0.16 Heywood (1996)<br />

hydroxide and 28.1<br />

magnesium<br />

hydroxide)<br />

Flocculated kaolin 0.75 �m 8.9–36.3 0.3–39 0.117–0.285 Heywood (1996)<br />

Deflocculated kaolin 0.75 �m 31.3–63.7 0.011–0.6 0.82–1.56 Heywood (1996)<br />

Magnesium hydroxide 5 �m 8.4–45.3 0.5–68 0.12–0.16 Heywood (1996)<br />

Pulverized fuel ash 38 �m 63–71.8 3.3–9.3 0.44–0.46 Heywood (1996)<br />

(PFA-P)<br />

Pulverized fuel ash 20 �m 70–74.4 2.12–9.02 0.48–0.57 Heywood (1996)<br />

(PFA-P)


3.26 CHAPTER THREE<br />

Shear stress<br />

(in units of pressure)<br />

1<br />

0.1<br />

0.01<br />

0.001<br />

0.0001<br />

slo<strong>pe</strong> = y/x<br />

Consider d�/dt = 700. Check �w = K(d�/dt) n .<br />

62.9 = 26.5 × 7000.132 This is close to the measured stress of 63.2 Pa. Therefore, the equation of this phosphate<br />

<strong>slurry</strong> is:<br />

�w = 26.5(d�/dt) 0.132<br />

The coefficient of rigidity is obtained as:<br />

Power Law Consistency Factor K<br />

Pa.s<br />

n<br />

/cm<br />

2<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0<br />

x<br />

0 1 10 100 1000 10,000<br />

Shear rate (1/sec)<br />

20<br />

clays<br />

magnetite<br />

40<br />

Volume Fraction of<br />

solids, C V<br />

y<br />

0<br />

n = y/x<br />

FIGURE 3-11 Plotting the rheology on a logarithmic scale to obtain the consistency factor<br />

“K” and the flow behavior index “n” of Pseudoplastics.<br />

Flow Behavior Index "n"<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0<br />

magnetite<br />

clays<br />

20 40<br />

K<br />

Volume Fraction of<br />

solids, CV<br />

FIGURE 3-12 Effect of volumetric concentration on the consistency factor “K” and the flow<br />

behavior index “n” of Pseudoplastics (after Aziz and Govier, 1972).<br />

n


at d�/dt = 700<br />

at d�/dt = 600.<br />

MECHANICS OF SUSPENSION OF SOLIDS IN LIQUIDS<br />

� = K(d�/dt) n–1<br />

� = 26.5(d�/dt) –0.878<br />

� = 26.5 × (700) –0.878<br />

� = 0.084 Pa · s<br />

� = 26.5 × 600 = 0.096 Pa · s<br />

3.27<br />

3-4-2-2 Pseudohomogeneous Pseudoplastics<br />

Pseudohomogeneous pseudoplastics behave similarly to their homogeneous counterparts.<br />

Clay sus<strong>pe</strong>nsions and magnetite-based slurries demonstrate an exponential relationship<br />

between n and C v as shown in Figure 3-12. The power law factor K has a more complex<br />

relationship with C v, as shown in Figure 3-12.<br />

Various equations have been derived to solve the power law factor of pseudoplastics.<br />

These equations are presented to help the reader appreciate the rheological constants that<br />

must be determined by testing, as will be described in Section 3-6.<br />

The Prandtl–Eyring equation is based on Dahlgreen’s (1958) discussion of the study<br />

conducted by Eyring and Prandtl on the kinetic theory of liquids:<br />

� = A sinh –1 [(d�/dt)/B] (3-46)<br />

where<br />

A and B = the rheological constants<br />

sinh = the hy<strong>pe</strong>rbolic function<br />

From Equation 3-44, the apparent viscosity is derived as<br />

�a = {A/(d�/dt)}{sinh –1 [(d�/dt)/B]} (3-47)<br />

The Ellis equation is more flexible but is an empirical equation and uses three rheological<br />

constants. Skelland (1967) demonstrates how the equation is based on the work of Ellis<br />

and Round and is explicit with res<strong>pe</strong>ct to the velocity gradient rather than the shear rate:<br />

(d�/dt) = (A0 + A1� (�–1) )�w (3-48)<br />

where A0, A1, and � are the rheological coefficients of the <strong>slurry</strong> material.<br />

The apparent viscosity is expressed as<br />

(�–1) �a = 1/(A0 + A1�w ) (3-49)<br />

When A1 = 0, the equation takes on a Newtonian form where A0 = 1/�.<br />

The equation reduces to the conventional power law equation with � = 1/n and A1 =<br />

(1/k) 1/n . When � > 1, the equation approaches a Newtonian flow at low shear stresses, and<br />

when � < 1, it tends to approach a Newtonian flow at high shear stress.<br />

The Cross equation (Cross, 1965) is a versatile equation that is based on measurements<br />

of viscosity, �0 at zero shear rate and �� at infinite shear rates.<br />

�� – �0 �a = �0 + ��<br />

(3-50)<br />

2/3<br />

1 + �(d�/dt)<br />

where � is a coefficient used to express to the shear stability of the mixture.<br />

This equation has been tested and has successfully predicted the behavior of a wide


3.28 CHAPTER THREE<br />

variety of pseudoplastic mixtures, such as sus<strong>pe</strong>nsions of limestone, non-aqueous polymer<br />

solutions, and nonaqueous pigment paste.<br />

3-4-3 Dilatant Slurries<br />

Dilatant fluids are time-inde<strong>pe</strong>ndent non-Newtonian fluids and are characterized by the<br />

following:<br />

� An infinitesimal shear stress is sufficient to initiate motion.<br />

� The rate of increase of shear stress with res<strong>pe</strong>ct to the velocity gradient increases as the<br />

velocity gradient increases.<br />

Dilatant fluids, therefore, use similar equations as pseudoplastic fluids. They are much<br />

less common than pseudoplastics. Dilatancy is observed under s<strong>pe</strong>cific conditions such as<br />

certain concentrations of solids, shear rates, and the sha<strong>pe</strong> of particles. Dilatancy is due to<br />

the shift, under shear action, of a close packing of particles to a more o<strong>pe</strong>n distribution in<br />

the liquid.<br />

Govier et al. (1957) observed the phenomena of dilatancy in sus<strong>pe</strong>nsions of magnetite,<br />

galena, and ferrosilicon in a range of particle sizes from 5 microns to 70 microns.<br />

It is observed that the slo<strong>pe</strong> of the shear stress versus the shear rate increases, particularly<br />

in the range of shear rates from 80 to 120 sec –1 . Metzener and Whitlock (1958) explained<br />

the phenomenon of dilatancy as follows.<br />

Two mechanisms account for the inflection and subsequent increase in the slo<strong>pe</strong> of<br />

the curve. Initially, the shear stress approaches a magnitude at which the size of flowing<br />

particles and aggregates is at a minimum and a Newtonian behavior develops (at<br />

the inflection of the curve). As the level of stress rises, the mixture expands volumetrically,<br />

and entire layers of particles start to slide or glide over each other. In the interim,<br />

the <strong>slurry</strong> acts as a pseudoplastic until the shear stress is high enough to cause dilatancy.<br />

The phenomenon of dilatancy is not easy to model. According to Metzener and Whitlock<br />

(1958), it is observed at volumetric concentration in excess of 27–30% and at shear<br />

rates in excess of 100 s –1 .<br />

3-4-4 Yield Pseudoplastic Slurries<br />

Yield pseudoplastic fluids are time-inde<strong>pe</strong>ndent non-Newtonian fluids and are characterized<br />

by the following:<br />

� An infinitesimal shear stress is sufficient to initiate motion.<br />

� The rate of increase of shear stress, with res<strong>pe</strong>ct to the velocity gradient, decreases as<br />

the velocity gradient increases.<br />

� A yield stress must be overcome at zero shear rate for motion to occur.<br />

Examples of yield pseudoplastics are shown in Table 3-11.<br />

Equation 3-44 is then modified to account for the yield stress as follows:<br />

�w – �0 = K[(d�/dt) n ] (3-51)<br />

Equation 3-51 is known as the Herschel–Buckley equation of yield pseudoplastics and<br />

is accepted by most <strong>slurry</strong> ex<strong>pe</strong>rts to describe the rheology of yield pseudoplastics with


MECHANICS OF SUSPENSION OF SOLIDS IN LIQUIDS<br />

TABLE 3-11 Examples of Yield Pseudoplastics<br />

Range of Angle of<br />

consistency flow<br />

Density, Yield stress coefficient K, behavior<br />

Slurry kg/m 3 � 0, Pa Ns n /m 2 index, n Reference<br />

3.29<br />

Sewage sludge 1024 1.268 0.214 0.613 Chilton and Stainsby (1998)<br />

Sewage sludge 1011 0.727 0.069 0.664 Chilton and Stainsby (1998)<br />

Sewage sludge 1013 2.827 0.047 0.806 Chilton and Stainsby (1998)<br />

Sewage sludge 1016 1.273 0.189 0.594 Chilton and Stainsby (1998)<br />

Kaolin <strong>slurry</strong> 1071 1.880 0.010 0.843 Chilton and Stainsby (1998)<br />

Kaolin <strong>slurry</strong> 1061 1.040 0.014 0.803 Chilton and Stainsby (1998)<br />

Kaolin <strong>slurry</strong> 1105 4.180 0.035 0.719 Chilton and Stainsby (1998)<br />

low to moderate concentration of solids. At high shear rates, certain complex phenomena<br />

such as dilatancy may develop. Certain bentonite clays develop a yield pseudoplastic rheology<br />

at 20% concentration by volume.<br />

Krusteva (1998) investigated the rheology of a number of inorganic waste slurries<br />

such as drilling fluids in <strong>pe</strong>troleum output, residue mineral materials in tailing ponds, filling<br />

of abandoned mine galleries, etc. In the case of clay containing industrial wastes, he<br />

indicated that colloidal forces of attraction or repulsion are ever present with Brownian<br />

forces and may cause thermodynamic instability. Waste materials such as blast furnace<br />

slag, fly ash, and material from mine filling exhibited various forms of a yield pseudoplastic<br />

rheology.<br />

The behavior of yield pseudoplastics can be expressed by the Carson model as described<br />

by Lapasin et al. (1998):<br />

�n = � n n<br />

0 +�� (d�/dt) (3-52)<br />

By binary system, Lapasin meant a mixture of two sizes of particles above the colloidal<br />

range and by ternary, three sizes. Alumina powders with average d50 diameters of 0.9<br />

�m, 1.4 �m, and 3.9 �m, and different s<strong>pe</strong>cific surface areas (8.23 m2 /cm3 , 5.74<br />

m2 /cm3 , and 2.65 m2 /cm3 ) were investigated. A dis<strong>pe</strong>rsing agent was used. Appreciable<br />

time-de<strong>pe</strong>ndent effects were only noticed at a concentration of the dis<strong>pe</strong>rsing agent below<br />

a critical value. Multicomponent sus<strong>pe</strong>nsions were found to have a viscosity that<br />

was de<strong>pe</strong>ndent on the total volume concentration of solids Cv and on the composition of<br />

the dis<strong>pe</strong>rsed phase expressed as a volume fraction. It was also de<strong>pe</strong>ndent on the shear<br />

rate of the mixture.<br />

Vlasak et al. (1998) investigated the addition of <strong>pe</strong>ptizing agents to kaolin–water mixtures.<br />

These mixtures were described as yield pseudoplastics that follow the<br />

Bulkley–Herschel rheological model (these will be discussed in Chapter 5). The addition<br />

of <strong>pe</strong>ptizing agents initially achieved a rapid drop of viscosity down to 8–10% of the original<br />

value up to an optimum concentration. As the concentration of the <strong>pe</strong>ptizing agent is<br />

increased beyond an optimum value, its effects are neutralized and the viscosity of the<br />

<strong>slurry</strong> increases again. Soda Water-GlassTM as a <strong>pe</strong>ptizing agent seemed to achieve the<br />

best reduction in viscosity when added at a concentration of 0.4%. The effect was a drastic<br />

drop of viscosity by 92% of its original value (without the <strong>pe</strong>ptizing agent). The optimum<br />

concentration of sodium carbonate, another <strong>pe</strong>ptizing agent, was 0.1%. The viscosity<br />

was reduced by 90%. These narrow bands of concentration of <strong>pe</strong>ptizing agents can<br />

effectively reduce the cost of hydro-transporting kaolin–water mixtures by reducing viscosity<br />

and therefore the coefficient of friction.


3.30 CHAPTER THREE<br />

3-5 TIME-DEPENDENT NON-NEWTONIAN<br />

MIXTURES<br />

Because crude oils and slurries of tar sands from certain Canadian mining projects develop<br />

a time-de<strong>pe</strong>ndent non-Newtonian behavior in cold tem<strong>pe</strong>ratures, a section of this chapter<br />

will pay attention to these complex thixotropic pro<strong>pe</strong>rties.<br />

In time-de<strong>pe</strong>ndent non-Newtonian flows, the structure of the mixture and the orientation<br />

of particles are sensitive to the shear rates. Due to structural changes and reorientation<br />

of particles at a given shear rate, the shear stress becomes time-de<strong>pe</strong>ndent as the particles<br />

realign themselves to the flow. In other words, the shear stress takes time to readjust<br />

to the prevailing shear rate. Some of these changes may be reversible when the rate of reformation<br />

is the same as the rate of decay. However, in the case of flows in which the deformation<br />

is extremely slow, the structural changes or particle reorientation may be irreversible<br />

(see Figure 3-13).<br />

3-5-1 Thixotropic Mixtures<br />

When the shear stress of a fluid decreases with the duration of shear strain, the fluid is<br />

called thixotropic. The change is then classified as reversible and structural decay is observed<br />

with time under constant shear rate. Certain thixotropic mixtures exhibit as<strong>pe</strong>cts of<br />

<strong>pe</strong>rmanent deformation and are called false thixotropic.<br />

When the rate of structural reformation exceeds the rate of decay under a constant sustained<br />

shear rate, the behavior is classified as rheo<strong>pe</strong>xy (or negative thixotropy).<br />

One typical example of a thixotropic mixture is a water sus<strong>pe</strong>nsion of bentonitic<br />

clays. These difficult slurries are produced by mud drilling associated with the use of<br />

positive displacement diaphragm or hose pumps. The reader may find throughout literature<br />

considerable discussion about “hysterisis.” This function is used to measure the<br />

behavior of the mixture by gradually increasing the shear rate and then by decreasing it<br />

back in steps. These curves are interesting but are of limited help to the designer of a<br />

pumping system.<br />

Shear Stress ( )<br />

Thixotropic<br />

Rheo<strong>pe</strong>ctic<br />

Rate of shear ( = du/dy)<br />

FIGURE 3-13 Rheology of time-de<strong>pe</strong>ndent fluids.


MECHANICS OF SUSPENSION OF SOLIDS IN LIQUIDS<br />

3.31<br />

Moore (1959) proposed expressing the complex behavior of a thixotropic fluid that<br />

does not possess a yield stress value in terms of six parameters:<br />

� = (�0 + c�)(d�/dt)<br />

d�/d� = a – �(a + bd�/dt)<br />

where<br />

� = duration of the shear for a time-de<strong>pe</strong>ndent fluid<br />

a, b, c, and �0 = materials constants<br />

� = a structural parameter that has two values (0 and 1) at the limits where<br />

the material is fully broken down or fully develo<strong>pe</strong>d<br />

Fredrickson (1970) discussed the modeling of thixotropic mixtures of sus<strong>pe</strong>nsions of<br />

solids in viscous liquids and proposed that rheological tests be conducted to measure four<br />

constants to understand the qualitative nature of the mixture.<br />

Ritter and Govier (1970) proposed representing the behavior of thixotropic fluids as<br />

follows:<br />

� The formation of structures, networks, or agglomerates is similar to a second-order<br />

chemical reaction.<br />

� The breakdown of the structure is similar to a series of consecutive first-order chemical<br />

reactions where formation is meant by behavior that is time-de<strong>pe</strong>ndent, whereas the<br />

breakdown occurs when the viscosity of the fluid acts as a Newtonian mixture that is<br />

inde<strong>pe</strong>ndent of both the shear rate and the duration of shear (Figure 3-14).<br />

Shear stress, +0.01, lb /ft<br />

-1 10<br />

8<br />

6<br />

4<br />

10<br />

4<br />

2<br />

2<br />

-2<br />

10<br />

Duration of<br />

shear, min<br />

0<br />

1<br />

10<br />

100<br />

100 1000<br />

-1<br />

Rate of Shear, d /dt + 10 in sec<br />

FIGURE 3-14 Rheology of Pembina crude oil at 44.5°F at constant duration of shear. (After<br />

Govier and Aziz, 1972.)


3.32 CHAPTER THREE<br />

Ritter and Govier (1970) therefore proposed to express the shear stress of the fluid in<br />

terms of structural stress � s and � �, a component of shearing stress due to the Newtonian<br />

component of the fluid:<br />

� = � s + � �<br />

(3-53)<br />

log� � = –KD� �log � – log K �s – �s� �s0 + �s� ��<br />

DR (3-54)<br />

where<br />

�s0, �s� = structural stresses at a given shear rate after zero and infinite duration of shear<br />

�s0 = �0 – �(d�/dt)<br />

�s� = �� – �(d�/dt)<br />

KD = a constant that is inde<strong>pe</strong>ndent of shear rate but is related to the first-order structural<br />

decay process and is expressed in the minutes –1 .<br />

KDR = a dimensionless measure of the interaction between the network or structure decay<br />

and the reestablishment processes<br />

The coefficient KDR is evaluated as<br />

�<br />

KDR = (3-55)<br />

where �s1 is measured after a lapse of 1 minute. In Equations 3-54 and 3-55, KDR, KD, �s0, �s1, and �s� are determined from rheology tests.<br />

Kherfellah and Bekkour (1998) examined the thixotropy of sus<strong>pe</strong>nsions of montmorillonite<br />

and bentonite clays. Montmorillonite clays are used as thickening agents for<br />

drilling fluids, paints, <strong>pe</strong>sticides, cosmetics, pharmaceuticals, etc. Commercial bentonite<br />

sus<strong>pe</strong>nsions exhibited thixotropic pro<strong>pe</strong>rties for concentrations higher than 6% by weight.<br />

Rheo<strong>pe</strong>ctic or negative thixotropic mixtures are not common in mining and will not be<br />

examined in this chapter.<br />

2 s0 – �s1�s� ��<br />

�s1�s� – � 2 � 2 (� s0/�s�) – �s �s0 – �s� s�<br />

3-6 DRAG COEFFICIENT OF SOLIDS<br />

SUSPENDED IN NON-NEWTONIAN FLOWS<br />

Some solids may be transported by highly viscous fluids in a non-Newtonian flow<br />

regime. One such example includes solids transported in the process of drilling a tunnel in<br />

a sandy soil rich with clay or bentonite. Other examples of solids sus<strong>pe</strong>nded in non-Newtonian<br />

flows are energy slurries, which are mixtures of fine coal and crude oils. In such<br />

circumstances, the drag coefficient of the coarse components is of interest.<br />

Brown (1991) reviewed the literature for settling of solids in non-Newtonian flows,<br />

but cautioned that the studies have been limited to single particles. Considerably more research<br />

is needed in this field.<br />

3-7 MEASUREMENT OF RHEOLOGY<br />

In the proceeding sections of this chapter, the concepts of Newtonian and non-Newtonian<br />

fluids were explored. Measuring the viscosity of a <strong>slurry</strong> mixture is recommended for ho-


MECHANICS OF SUSPENSION OF SOLIDS IN LIQUIDS<br />

mogeneous flows, mixtures with a high concentration of particles, and for fibrous and<br />

flocculated slurries.<br />

Subsieve particles are defined as particles with an average diameter smaller than<br />

35–70 �m (de<strong>pe</strong>nding on whose reference book you consult). Slurry flows with subsieve<br />

particles at a relatively high concentration by volume (C v � 30%) are strongly rheologyde<strong>pe</strong>ndent.<br />

Heterogeneous flows, flows without subsieve particles, or flows with subsieve<br />

particles at a very low concentrations, are not governed by the rheology of the <strong>slurry</strong>.<br />

Flocculation or the addition of flocculates in the process of mixing slurries tends to result<br />

in non-Newtonian rheology.<br />

Rheology in simple layman’s terms is the relationship between the shear stress and the<br />

shear rate of the <strong>slurry</strong> under laminar flow conditions. Although this relationship extends<br />

to transitional and turbulent flows, most tests are conducted in a laminar regime, often in<br />

tubes or between parallel plates.<br />

3-7-1 The Capillary-Tube Viscometer<br />

The purpose of the capillary tube viscometer is to measure the rheology of a laminar flow<br />

under controlled velocity conditions. Tubes are used in a range of diameters from 0.8–12<br />

mm (1/32–1/2 in). The length of the tube is accurately cut to account for entrance effects<br />

and end effects. Typically, the length may be as much as 1000 times the inner diameter.<br />

The capillary tube viscometer is used to plot the average rate versus the shear stress at<br />

the wall of the tube. This is called the pseudoshear diagram, as defined by the<br />

Mooney–Rabinovitch equation:<br />

(du/dr) w = �0.75 + 0.2 8<br />

d[ln(8V/Di)] � ���<br />

(3-56)<br />

Di<br />

d[ln(�P/4Li)]<br />

where<br />

(du/dr) w = rate of shear at the wall<br />

�P = pressure drop due to friction over a length Li of pi<strong>pe</strong> of inner diameter Di V = average velocity of the flow<br />

d = derivative<br />

The data is then plotted on a logarithmic scale as <strong>pe</strong>r Figure 3-15.<br />

The use of capillary-like viscometers is complicated by the “effective slip” of non-<br />

Newtonian fluid-sus<strong>pe</strong>nded material, which tends to move away from the wall, leaving an<br />

attached layer of liquid. The result is a reduction in the measurements of effective viscosity.<br />

Therefore, it is often recommended to conduct such tests in a number of tubes of different<br />

diameters.<br />

Measuring the pressure loss between two points well away from the entrance and end<br />

effects gives the shear stress at the wall as:<br />

�w = Ri�P/(2Li) (3-57)<br />

By considering that the velocity profile at a height y above the wall is a function of the<br />

shear stress we obtain<br />

–(du/dy) w = f (�)<br />

It may be possible to establish a relationship between the flow rate Q and the shear stress<br />

� as<br />

Q<br />

� �R 3<br />

1<br />

= � �w �3 � w 0<br />

3.33<br />

� 2 f (�)d� (3-58)


3.34 CHAPTER THREE<br />

8V<br />

D<br />

shear rate<br />

For a Newtonian flow:<br />

or � = � w/(8V/D i).<br />

For a Bingham flow:<br />

for � > � 0, where � 0 is the yield stress.<br />

The velocity profile is expressed as<br />

2V �w = � = � (3-59)<br />

Di 4�<br />

� = �(du/dr) w + � 0<br />

= = � �2 2V � �w (� – �0)d� (3-60)<br />

By integration of this equation and by multiplying by 4, the shear rate is derived as<br />

8V<br />

� DI<br />

100<br />

10<br />

1.0<br />

Q<br />

� �R 3<br />

0<br />

0<br />

� Di<br />

� w<br />

� �<br />

water<br />

Q<br />

� �R 3<br />

increasing tube diameter<br />

Shear Stress<br />

� � w 3<br />

4<br />

�<br />

3<br />

� 0<br />

� �w<br />

= � 1 – � � + � �� (3-61)<br />

Equation 3-61 is called the Buckingham equation. This equation cannot be solved<br />

without long iterations. Many engineers prefer to simplify the Buckingham equation by<br />

ignoring the term (� 0/� w) 4 , as this term is of negligible magnitude compared with the other<br />

terms:<br />

� �0<br />

1.0 10<br />

1<br />

�<br />

3<br />

D<br />

2<br />

D 1<br />

D 3<br />

4 � 0<br />

�4 � w<br />

D 4<br />

D P<br />

4 L<br />

FIGURE 3-15 Pseudoshear diagram of a non-Newtonian mixture tested in a capillary tube<br />

rheometer.


MECHANICS OF SUSPENSION OF SOLIDS IN LIQUIDS<br />

�w � 8V�/Di + 4/3�0 The modified equation is plotted in Figure 3-16.<br />

(3-62)<br />

For a pseudoplastic <strong>slurry</strong> or power law fluid, the shear stress is expressed by Equation<br />

3-43. By analogy with the method develo<strong>pe</strong>d for a Bingham flow in a tube, the following<br />

equation is expressed:<br />

= = � � 2 (�/K) 1/n Q 2V<br />

�3 �R<br />

1<br />

�3 �w<br />

�<br />

d� (3-63)<br />

� Di<br />

or<br />

= � �w (3-64)<br />

0<br />

which once integrated is expressed as<br />

= � � (3-65)<br />

The effective viscosity is expressed as<br />

�e = �w/(8V/Di) = K(8V/Di) (n–1) [4n/(3n + 1)] n �<br />

1/n<br />

2V n � w<br />

� � �1/n Di 3n + 1 K<br />

(3-66)<br />

Unfortunately, Equation 3-66 is of no value when n < 1.0, which is the case for many<br />

power law slurries. It would mean that as the shear rate increases, the effective viscosity<br />

decreases to zero. This is contradictory to nature. For power law exponents smaller than<br />

1.0, alternative equipment should be used to measure rheology.<br />

It is tricky to avoid errors with the use of capillary effect viscometers. A particular<br />

source of errors is the end effect. At the entrance exit of the tube, contraction and expansion<br />

of the flow cause additional pressure losses.<br />

(3+1/n)<br />

Q<br />

� ��<br />

3<br />

1/n<br />

�R (3 + 1/n)K<br />

w<br />

Shear Stress<br />

Velocity profile<br />

w<br />

2 r 0<br />

� �0<br />

shear rate<br />

FIGURE 3-16 Pseudoshear diagram for a Bingham plastic.<br />

dV<br />

�<br />

dy<br />

dU<br />

dy<br />

3.35


3.36 CHAPTER THREE<br />

3-7-2 The Coaxial Cylinder Rotary Viscometer<br />

A more practical instrument to use when measuring rheology is the coaxial cylinder viscometer.<br />

In basic terms, it is a device used to measure the resistance or torque when rotating<br />

a cylinder in a viscous fluid (Figure 13-17). The moment of inertia in the cylinder is<br />

established by the manufacturer.<br />

The torque is due to the force the fluid exerts tangentially to the outside surface of the<br />

cylinder:<br />

T = 2�R0h�wR0 (3.67)<br />

where<br />

T = (surface area) (shear stress) (radius)<br />

R0 = outside radius of the rotating cylinder<br />

h = height of the cylinder<br />

�w = shear stress at the wall<br />

The shear stress at any radius r in the fluid can be expressed as<br />

T du<br />

�w = � = ��<br />

2 2�r h dy<br />

If the liquid is rotating at an angular velocity �, then<br />

(du/dy) w = –rd�/dr<br />

r<br />

R c<br />

scale to measure torque<br />

rotation of bob at<br />

s<strong>pe</strong>ed<br />

R<br />

FIGURE 3-17 The rotating concentric viscometer.<br />

0<br />

<strong>slurry</strong><br />

(3.68)


and<br />

MECHANICS OF SUSPENSION OF SOLIDS IN LIQUIDS<br />

� = –�rd�/dr<br />

d�<br />

� dr<br />

=<br />

� �<br />

d� = � 0<br />

Rc R0 –T<br />

� 2�h�r<br />

–T<br />

� dr 3 2�h�r<br />

3.37<br />

or<br />

� = � – � (3.69)<br />

where Rc is the radius of the outside cylinder.<br />

2 This is known as the Margulus equation. It is obvious that R 0 can be related to the moment<br />

of inertia Ik of the rotating bob cup.<br />

Since for a Bingham <strong>slurry</strong>, the rate of shear is expressed as du/dr = (� – �0); the Margulus<br />

equation can be demonstrated as<br />

� = � – � – ln� � (3.70)<br />

This equation is known as the Reiner–Rivlin equation.<br />

For a Pseudoplastic:<br />

2 1/n � = n[T/(2�R 0hK)] [1 – (R0/Rc) 2/n T 1 1<br />

� � � 2 2<br />

4�h� R 0 R c<br />

T 1 1 �0 Rc � � � � � 2 2<br />

4�h� R 0 R c � R0<br />

] (3.71)<br />

At the wall:<br />

2 �w = T/(2�R b h) (3.72)<br />

A plot of log �w versus log � can be constructed. The slo<strong>pe</strong> gives the flow index n and, by<br />

substituting Equation 3-45, the value of K can be calculated.<br />

Heywood (1991) discussed errors with the use of rotating viscometers. Particular<br />

sources of errors are the end effects from both cylinders and the possible deformation of<br />

the laminar layer under the effect of high rotational s<strong>pe</strong>ed. Heywood recommended the<br />

use of cylinders with a long length to diameter ratio. Wall slip effects can be detected by<br />

using cylinders of different radius but same length. The vendors of rheometers publish<br />

equations to correct for wall slip and end effects.<br />

One important problem about the use of rheometers is that they do not distinguish between<br />

Bingham and Carson slurries. This can lead to grave mistakes in the design of a<br />

pi<strong>pe</strong>line. Certain slurries have a course of fractions that could also precipitate during a<br />

rheometer test. Unfortunately, this would give false readings. When there is doubt, the<br />

safest approach is to conduct a pro<strong>pe</strong>r pump test in a loop.<br />

Whorlow (1992) published a book on rheological techniques that includes dynamic<br />

tests and wave propagation tests. In the ap<strong>pe</strong>ndix, he listed a number of rheological investigation<br />

equipment manufacturers. Some of the techniques apply more to polymers and<br />

are not relevant to our discussion. Dynamic vibration tests have been extended to fresh<br />

concrete (Teixera et al., 1998). Concord and Tassin (1998) described a method to use<br />

rheo-optics for the study of thixotropy in synthetic clay sus<strong>pe</strong>nsions. A rheometer optical<br />

analyzer was used on laponite, a synthetic hectorite clay. Laponite was mixed with water<br />

and tests were conducted at various intervals for up to 100 days. Rheo-optics seems to be


3.38 CHAPTER THREE<br />

FIGURE 3-18 Stresstech rheometer, courtesy of ATS Reho<strong>systems</strong>. The rheometer was develo<strong>pe</strong>d<br />

for the pharmaceutical and cosmetics industries, where materials consistency may<br />

vary from fluid to solid.<br />

a new technique based on the ability of solids to reorient themselves by applying to them<br />

a negative electrical charge.<br />

3-8 CONCLUSION<br />

In this chapter, it was demonstrated that mixtures of solids and liquids are complex <strong>systems</strong>.<br />

The size of the particles, the diameter of the pi<strong>pe</strong>, the interaction with other particles,<br />

the viscosity of the carrier, and the tem<strong>pe</strong>rature of the flow all interact to yield Newtonian<br />

or non-Newtonian flows.<br />

In the next three chapters, the principles discussed in the present chapter will be applied<br />

to calculate the velocity of deposition, the critical velocity, the stratification ratio,<br />

and the friction loss in closed and o<strong>pe</strong>n conduits for heterogeneous and homogeneous<br />

mixtures.<br />

3-9 NOMENCLATURE<br />

a The longest axis of a particle in Albertson’s model<br />

A Parameter used to express viscosity of non-Newtonian flows


MECHANICS OF SUSPENSION OF SOLIDS IN LIQUIDS<br />

3.39<br />

A0 Coefficient<br />

A1 Coefficient<br />

b Axis of a particle in Albertson’s model<br />

B Parameter used to express viscosity of non-Newtonian flows<br />

c The shortest axis of a particle in Albertson’s model<br />

C Parameter used to express viscosity of non-Newtonian flows<br />

CD Drag coefficient of an object moving in a fluid<br />

CDo Profile drag coefficient of an object moving in a fluid<br />

CL Lift coefficient of an object moving in a fluid<br />

CN CS Cv Cv� Cw da Coefficient based on equivalent Ren Coefficient based on equivalent Ren Concentration by volume of the solid particles in <strong>pe</strong>rcent<br />

Maximum packing concentration of solids<br />

Concentration by weight of the solid particles in <strong>pe</strong>rcent<br />

Diameter of a sphere with a surface area equal to the surface area of the irregularly<br />

sha<strong>pe</strong>d particle<br />

dapp Apparent particle diameter<br />

df Apparent flocculant diameter<br />

dg Sphere diameter<br />

dn Diameter of a sphere with a volume equal to the volume of the irregularly<br />

sha<strong>pe</strong>d particle in Albertson’s model<br />

d� Particle diameter<br />

D Drag force<br />

Di Tube or pi<strong>pe</strong> inner diameter<br />

E Factor between Albertson and Clift sha<strong>pe</strong> factors<br />

f( ) Function of<br />

FBF Buoyancy force<br />

Fw Wall effect correction factor for free-fall s<strong>pe</strong>ed of a particle<br />

g Acceleration due to gravity (9.78–9.81 m/s2 )<br />

gc Conversion factor, 32.2ft/s2 if U.S. units between lbms and slugs<br />

h Height of the cylinder<br />

Ik Moment of inertia<br />

K Consistency index or power law coefficient for a pseudoplastic<br />

KD A constant that is inde<strong>pe</strong>ndent of shear rate but is related to the first-order<br />

structural decay process and is express in minutes –1<br />

KDR A dimensionless measure of the interaction between the network or structure<br />

decay and the reestablishment processes<br />

Kt Coefficient for terminal velocity<br />

Kz Kozney constant<br />

K1, K2, K3 Coefficients<br />

ln natural logarithm<br />

L Lift force<br />

Lc Characteristic length<br />

LI Length of pi<strong>pe</strong> or tube<br />

n Flow behavior index, or exponent for a pseudoplastic (


3.40 CHAPTER THREE<br />

Ren Reynolds Number of a particle based on dn Rep Reynolds Number of a sphere particle based on its diameter<br />

Ri Inner radius of a pi<strong>pe</strong> or tube<br />

R0 Radius of the bob in the coaxial cylinder rotary viscometer<br />

sp The surface area <strong>pe</strong>r unit volume of a sphere of equivalent dimensions or<br />

6/dg, also called s<strong>pe</strong>cific surface of a particle<br />

Sf Front area of a particle orthogonal to the direction of flow<br />

Sw Surface area of a wing along the direction of flight<br />

T Applied torque for the cylinder rotary viscometer<br />

Ta Absolute tem<strong>pe</strong>rature<br />

V Average velocity of the flow<br />

V0 Terminal velocity at very low volume concentration of solids<br />

Vc Terminal velocity at given volume concentration of solids<br />

Vt The terminal (or free settling) s<strong>pe</strong>ed<br />

W Weight<br />

� the ratio of immobilized dis<strong>pe</strong>rsing fluid to the volume of solids related approximately<br />

to the particle and floc apparent diameter<br />

� A coefficient used to express to the shear stability of a pseudoplastic mixture<br />

� Concentration by volume in decimal points<br />

� Shear strain<br />

d�/dt Wall shear rate or rate of shear strain with res<strong>pe</strong>ct to time<br />

� Coefficient of rigidity of a non-Newtonian fluid, also called Bingham viscosity<br />

� A structural parameter for thixotropic fluids, which do not possess a yield<br />

stress value<br />

� Carrier liquid absolute viscosity<br />

�a Apparent viscosity of a pseudoplastic fluid<br />

�e Effective viscosity<br />

�0 Apparent viscosity of a pseudoplastic fluid at zero shear rate<br />

�� Bingham plastic limiting viscosity, or apparent viscosity of a pseudoplastic<br />

fluid at very high shear rate<br />

� Pythagoras number (ratio of circumference of a circle to its diameter)<br />

� Duration of the shear for a time-de<strong>pe</strong>ndent fluid<br />

� Density<br />

� Shear stress at a height y or at a radius r<br />

�0 Yield stress for a Bingham plastic or yield pseudoplastic<br />

�s Structural stress of a thixotropic fluid<br />

�w Wall shear stress<br />

� Kinematic viscosity<br />

� Angular velocity of particle<br />

� Angular velocity of complete system<br />

� The logarithmic standard deviation<br />

�A Albertson sha<strong>pe</strong> factor<br />

�c Clift sha<strong>pe</strong> factor<br />

Thomas sha<strong>pe</strong> factor<br />

� T<br />

Subscripts<br />

g Equivalent sphere<br />

L Liquid<br />

m Mixture<br />

p Particle<br />

s Solids


3–10 REFERENCES<br />

MECHANICS OF SUSPENSION OF SOLIDS IN LIQUIDS<br />

3.41<br />

Albertson, M. L. 1953. Effects of sha<strong>pe</strong> on the fall velocity of gravel particles. Pa<strong>pe</strong>r read at the 5th<br />

Iowa Hydraulic Conference, Iowa University, Iowa City, Iowa.<br />

Allen, H. S. 1900. The motion of a sphere in a viscous fluid. Phil. Mag., 50, 323–338, 519–534.<br />

Boger, D. V., and Q. D. Nguyen. 1987. The Flow Pro<strong>pe</strong>rties of Weipa #3 and #4 Plant Tailings. Internal<br />

study conducted by Comalco Aluminium Ltd, Weipa, Australia, quoted in Darby, R., R.<br />

Mun, and D. V. Boger. 1992. Predict Friction Loss in Slurry Pi<strong>pe</strong>s. Chem. Engineering, 99, 9<br />

(September), 117–211.<br />

Brown, G. G. 1950. Unit O<strong>pe</strong>rations. New York: Wiley.<br />

Brown, N. P. 1991. The settling behavior of particles in fluids. In Slurry Handling, Edited by N. P.<br />

Brown and N. I. Heywood. New York: Elsevier Applied Sciences.<br />

Caldwell, D. H., and H. E. Babitt. 1941. Flow of muds, sludge and sus<strong>pe</strong>nsions in circular pi<strong>pe</strong>. Am.<br />

Inst. Chem. Engrs. Trans., 37, 2 (April 25), 237–266.<br />

Caldwell, D. H., and H. E. Babitt. 1942. Pi<strong>pe</strong>line flow of solids in sus<strong>pe</strong>nsion. Symp. Am. Soc. of Civ.<br />

Eng. Proc., 68, 3 (March), 480–482.<br />

Cheng, D. C. H., and W. Whitaker. 1972. Applications of the Warren Spring Laboratory pi<strong>pe</strong>line design<br />

method to settling sus<strong>pe</strong>nsion. Pa<strong>pe</strong>r read at the 2nd Annual Hydrotransport Conference,<br />

Bedford, England.<br />

Chilton, R. A., and R. Stainsby. 1998. Pressure loss equations for laminar and turbulent non-Newtonian<br />

pi<strong>pe</strong> flow. Journal of Hydraulic Engineering, 124, 5 (May), 522–529.<br />

Clift, R., J. R. Grace, and M. E. Weber. 1978. Bubbles, Drops and Particles. New York: Academic<br />

Press.<br />

Concord, S., and J. F. Tassin. 1998. Rheoptical study of thixotropy in synthetic clay sus<strong>pe</strong>nsions. In<br />

Proceedings of the Fifth Euro<strong>pe</strong>an Rheology Conference. Ljubljana, Slovenia: University of<br />

Ljubljana.<br />

Cross, M. M. 1965. Rheology of non-Newtonian fluids—New flow equation for pseudoplastic <strong>systems</strong>.<br />

Journal of Colloid Science, 20, 417.<br />

Dahlgreen, S. E. 1958. Eyring model of flow applied to thixotropic equilibrium. Journal of Colloid<br />

Science, 13 (April), 151–158.<br />

Dedegil, M. Y. 1987. Drag coefficient and settling velocity of particles in non-Newtonian sus<strong>pe</strong>nsions.<br />

Journal of Fluids Engineering, 109 (September), 319–323.<br />

Dick, R. I., and B. B. Ewing. 1967. Rheology of activated sludge. Journal of Water Pollution Control<br />

Federation, 39, 543.<br />

Dahlgreen, S. E. 1958. Eyring model of flow applied to thixotropic equilibrium. Journal of Colloid<br />

Science, 13 (April), 151–158.<br />

Fredrickson, A. G. 1970. A model for the thixotropy of sus<strong>pe</strong>nsions. American Inst. of Chem. Eng.<br />

Journal, 16, 436.<br />

Gay E. D., P. A. Nelson, and W. P. Armstrong. 1969. Flow pro<strong>pe</strong>rties of sus<strong>pe</strong>nsions with high<br />

solids concentration. American Inst. of Chem. Eng. Journal, 15, 6, 815–822.<br />

Goodrich and Porter. 1967.<br />

Govier, G. W., C. A. Shook, and E. O. Lilge. 1957. Rheological Pro<strong>pe</strong>rties of water sus<strong>pe</strong>nsions of<br />

finely subdivided magnetite, galena, and ferrosilicon. Trans. Can. Inst. Mining Met., 60, 157.<br />

Govier, G. W., and K. Aziz. 1972. The Flow of Complex Mixtures in Pi<strong>pe</strong>s. New York: Van Nostrand<br />

Reinhold.<br />

Hedstrom, B. O. A. 1952. Flow of plastic materials in pi<strong>pe</strong>s. Ind. Eng. Chem., 33, 651–656.<br />

Herbrich, J. 1968. Deep ocean mineral recovery. Pa<strong>pe</strong>r read at the World Dredging Conference II,<br />

Rotterdam, the Netherlands.<br />

Herbrich, J. 1991. Handbook of Dredging Engineering. New York: McGraw-Hill.<br />

Heywood, N. I. 1991. Rheological characterisation of non-settling slurries. In Slurry Handling, Edited<br />

by N. P. Brown and N. I. Heywood. New York: Elsevier Applied Sciences.<br />

Heywood, N. I. 1996. The <strong>pe</strong>rformance of commercially available Coriolis mass flowmeters applied<br />

to industrial slurries. Pa<strong>pe</strong>r read at the 13th International Hydrotransport Symposium on<br />

Slurry Handling and Pi<strong>pe</strong>line Transport. Johannesburg, South Africa. Cranfield, UK: BHRA<br />

Group.<br />

Inter-Agency Committee on Water Resources. 1958. Report 12. Internal report by the Subcommittee<br />

on Sedimentation, Minneapolis, Minnesota.


3.42 CHAPTER THREE<br />

Kearsey, H. A., and L. E. Gill. 1963. Study of sedimentation of flocculated thorium slurries using<br />

gamma ray technique. Trans. Inst. Chem. Engrs., 41, 296.<br />

Kherfellah, N., and K. Bekkour. 1998. Rheological characteristics of clay sus<strong>pe</strong>nsions. In Proceedings<br />

of the Fifth Euro<strong>pe</strong>an Rheology Conference. Ljubljana, Slovenia: University of Ljubljana.<br />

Krusteva, E. 1998. Viscosmetric and pi<strong>pe</strong> flow of inorganic waste slurries. In Proceedings of the<br />

Fifth Euro<strong>pe</strong>an Rheology Conference. Ljubljana, Slovenia: University of Ljubljana.<br />

Lapassin, R., S. Pricl, and M. Stoffa. 1998. Viscosity of aqueous sus<strong>pe</strong>nsions of binary and ternary<br />

alumina mixtures. In Proceedings of the Fifth Euro<strong>pe</strong>an Rheology Conference. Ljubljana,<br />

Slovenia: University of Ljubljana.<br />

Metzner, A. B., and M. Whitlock. 1958. Flow behavior of concentrated (dilatant) sus<strong>pe</strong>nsions. Trans.<br />

Soc. Rheology, 2, 239–254.<br />

Moore, F. 1959. Rheology of Ceramic Slips and Bodies. British Ceramic Society Transactions, 58,<br />

470.<br />

Mun, R. 1988. The Pi<strong>pe</strong>line Transportation of Sus<strong>pe</strong>nsions with a Yield Stress. Master’s Thesis, University<br />

of Melbourne, Australia.<br />

Parzonka, W. 1964. Determination of the maximum concentration of homogeneous mixtures (in<br />

French). Journal of the French Academy of Science, 259, 2073.<br />

Pil<strong>pe</strong>l, N. 1965. Flow pro<strong>pe</strong>rties of non-cohesive powders. Chemical Process Eng. 46, 4, 167–179.<br />

Prokunin, A. N. 1998. Particle-wall interaction in liquids with different rheology. In Proceedings of<br />

the Fifth Euro<strong>pe</strong>an Rheology Conference. Ljubljana, Slovenia: University of Ljubljana.<br />

Richards, R. H. 1908. Velocity of Galena and Quartz Falling in Water. Trans AIME, 38, 230–234.<br />

Ritter, R. A., and G. W. Govier. 1970. The development and evaluation of a theory of thixotropic behavior.<br />

Can. Journal Chem. Eng., 48, 505.<br />

Rubey, W. W. 1933. Settling velocities of gravel, sand and silt particles. Amer. Journal of Science,<br />

25, 148, 325–338.<br />

Skelland, A. H. P. 1967. Non-Newtonian Flow and Heat Transfer. New York: Wiley.<br />

Teixeira, M. A. O. M., R. J. M. Craik, and P. F. G. Banfill. 1998. The effect of wave forms on the vibrational<br />

processing of fresh concrete. In Proceedings of the Fifth Euro<strong>pe</strong>an Rheology Conference.<br />

Ljubljana, Slovenia: University of Ljubljana.<br />

Thomas, D. G. 1961. Transport characteristics of sus<strong>pe</strong>nsions: Part II. Minimum transport velocity<br />

for flocculated sus<strong>pe</strong>nsions in horizontal pi<strong>pe</strong>s. AIChE Journal, 7 (September), 423–430.<br />

Thomas, D. G. 1963. Transport characteristics of sus<strong>pe</strong>nsions. Ch. E. Journal, 9, 310.<br />

Thomas, A. D. 1981. Slurry pi<strong>pe</strong>line rheology. Pa<strong>pe</strong>r presented at the National Conference on Rheology.<br />

Second Annual Conference of the British Society of Rheology, Australian Branch, University<br />

of Sydney, Australia.<br />

Turton, R., and O. Levenspiel. 1986. A short note on drag correlation for spheres. Powder Technology<br />

Journal, 47, 83.<br />

Valentik, L., and R. L. Whitemore. 1965. Terminal velocity of spheres in Bingham plastics. British<br />

Journal of Applied Phys., 16, 1197.<br />

Vlasak, P., Z. Chara, and P. Stern. 1998. The effect of additives on flow behaviour of kaolin–water<br />

mixtures. In Proceedings of the Fifth Euro<strong>pe</strong>an Rheology Conference. Ljubljana, Slovenia:<br />

University of Ljubljana.<br />

Wasp, E. J., J. P. Kenny, and R. L. Gandhi. 1977. Solid-Liquid Flow—Slurry Pi<strong>pe</strong>line Transportation.<br />

Trans-Tech Publications.<br />

Wells, P. J. 1991. Pumping non-Newtonian slurries. Technical Bulletin 14. Sydney, Australia: Warman<br />

International.<br />

Whorlow, R. W. 1992. Rheological Techniques, 2d. ed. New York: Ellis Horwood.<br />

Wilson, K. C., G. R. Addie, and R. Clift. 1992. Slurry Transport Using Centrifugal Pumps. New<br />

York: Elsevier Applied Sciences.<br />

Worster, R. C., and D. E. Denny. 1955. Hydraulic transport of solid materials in pi<strong>pe</strong>s. Proceedings<br />

of the Institute of Mechanical Engineers (UK), 38, 230–234.<br />

Further Reading:<br />

Caldwell, D. H., and H. E. Babitt. 1942. Pi<strong>pe</strong>line flow of solids in sus<strong>pe</strong>nsion. Symp. Am. Soc. of Civ.<br />

Eng. Proc., 68, 3 (March), 480–482.<br />

Goodrich, J. E., and R. S. Porter. 1967. Rheological interpretation of torque—Rheometer data. Polymer<br />

Eng & Science, 7 (January), 45–51.


MECHANICS OF SUSPENSION OF SOLIDS IN LIQUIDS<br />

3.43<br />

Lazerus, J. H., and P. T. Slatter. 1988. A method for the rheological characterization of tube viscometer<br />

data. Journal of Pi<strong>pe</strong>lines, 7, 165–176.<br />

Thomas, D. G. 1960. Heat and momentum transport characteristics of non-Newtonian aqueous thorium<br />

oxide. AIChE Journal, 7, 431.<br />

Wilson, K. C. 1991. Pi<strong>pe</strong>line design for settling slurries. In Slurry Handling. Edited by N. P. Brown,<br />

and N. I. Heywood. New York: Elsevier Applied Sciences.<br />

Wilson, K. C. 1991. Slurry transport in flumes. In Slurry Handling. Edited by N. P. Brown, and N. I.<br />

Heywood. New York: Elsevier Applied Sciences.


APPENDIX A<br />

SPECIFIC GRAVITY AND<br />

HARDNESS OF MINERALS<br />

The following abbreviations are used in the table below<br />

A Amphibole group M Mica group<br />

B Bauxite component O Orthoclase<br />

C Clay or clay-like Ov Olivine group<br />

D Diopside series P Pyroxene<br />

E Enstatite group R Rare earth group<br />

F Feldspar group S Spinel group<br />

Fp Feldspathoid group Sc Scapolite series<br />

G Garnet group W Wolframite<br />

H Hornblende Z Zeolite group<br />

J Jamesonite<br />

Name Description or composition S<strong>pe</strong>cific gravity Mohr hardness<br />

Acanthite Ag2S 7.2–7.3 2–2.5<br />

Achroite Colorless tourmaline<br />

Acmite NaFe(SiO3) 2 3.4–3.6 6–6.5<br />

Actinolite Ca2(MgFe) 5(Si8O22)(OH) 2 3.0–3.2 5–6<br />

Adularia Clear orthoclase<br />

Aergite Acmite, aegrinine<br />

Agate Banded chalcedony<br />

Alabandite MnS 4.03.4–4<br />

Alabaster Fine-grained gypsum<br />

Albite Na(AlSi3O8) Alexandrite Chrysoberyl, gemstone<br />

Allanite (Ce,Ca,Y)(Al,Fe) 3(SiO4) 3(OH) 3.5–4.2 5.5–6<br />

Allemontite AsSb 5.8–6.2 3–4<br />

Allophane Al2O3·SiO2·nH2O 1.8–1.9 3<br />

Almandite Fe3Al2(SiO4) 3, red 4.25 7<br />

Altaite PbTe 8.16 3<br />

Alunite KAl3(SO4) 2(OH) 6 2.6–2.8 4<br />

Amazonstone Green microcline<br />

Amblygonite (Li,Na)AlPO4(F,OH) 3.0–3.1 6<br />

Amethyst Purple quartz<br />

A.1


A.2 APPENDIX A<br />

Name Description or composition S<strong>pe</strong>cific gravity Mohr hardness<br />

Amphibole A group of minerals<br />

Analcime Na(AlSi2O6)·H2O 2.27 5–5.5<br />

Anastase TiO2 3.9 5.5–6<br />

Anauxite Silicon-rich kaolinite<br />

Andalusite Al2SiO5 3.1–3.2 7.5<br />

Andesine Ab70An30—Ab50An50 2.69 6<br />

Andradite Ca3Fe2(SiO4) 3 3.75 7<br />

Anglesite PbSO4 6.2–6.4 3<br />

Anhydrite CaSO4 2.8–3.0 3–3.5<br />

Ankerite Ca(Fe,Mg,Mn)(CO3) 2 2.9–3.0 3.5<br />

Annabergite (Ni,Co) 3(AsO4) 2·8H2O 3.0 3–3.5<br />

Anorthite CaAl2Si2O8 2.76 6<br />

Anorthoclase (Na,K)AlSi3O8 2.58 6<br />

Antigorite Ser<strong>pe</strong>ntine<br />

Antimony Sb 6.7 3<br />

Anterite Cu3SO4(OH) 4 3.9 3.5–4<br />

Apatite Ca5(PO4,CO3) 3(F,OH,Cl) 3.1–3.2 5<br />

Apophylite KCa4Si8O20(F,OH)·8H2O 2.3–2.4 4.5–5<br />

Aquamarine Green-blue beryl, gemstone<br />

Aragonite CaCO3 2.95 3.5–4<br />

Arfvedsonite Na2-3(Fe,Mg,Al) 5Si8O22(OH) 2 3.45 6<br />

Argentite Ag2S 7.3 2–2.5<br />

Arsenic As 5.7 3.5<br />

Arsenopyrite FeAsS 5.9–6.2 5.5–6<br />

Asbestos A group of minerals<br />

Altacamite Cu2Cl(OH) 3 3.7–3.8 3–3.5<br />

Augite (Ca,Na)(Mg,Fe,Al)(Si,Al) 2(O) 6 3.2–3.4 5–6<br />

Aurichalcite (Zn,Cu) 5(CO3) 2(OH) 6 3.2–3.7 2<br />

Autunite Ca(UO2) 2(PO4) 2·10H2O 3.1–3.2 2–2.5<br />

Awaruite FeNi2 7.7–8.1 5<br />

Axinite (Ca,Mn,Fe) 3Al2BSi4O15(OH) 3.2–3.4 6.5–7<br />

Azurite Cu3(CO3) 2(OH) 2 3.77 3.5–4<br />

Balas ruby Red Spinel, gemstone<br />

Barite BaSO4 4.5 3–3.5<br />

Barytes Barite<br />

Bastnaesite (Ce,La)(CO3)(F,OH) 4.9–5.2 4–4.5<br />

Bauxite Aluminum hydroxide mixture<br />

Beidellite Al8(Si4O10) 3(OH) 12·12H2O 2.6 1.5<br />

Bentonite Montmorillonite clay<br />

Beryl Be3Al2(Si6O18) 2.7–2.8 7.5–8<br />

Biotite K(Mg,Fe2+ )3(Al,Fe3+ )Si3O10(OH) 12 2.8–3.2 2.5–3<br />

Bismite Bi2O3 8 4.5<br />

Bismuth Bi 9.8 2–2.5<br />

Black Jack Sphalerite<br />

Blende Sphalerite<br />

Bloodstone Heliotro<strong>pe</strong><br />

Blue vitriol Chalcanthite<br />

Boehmite AlO(OH) 3.0–3.1<br />

Boracite Mg3B7O13Cl 2.9–3.0 7


SPECIFIC GRAVITY AND HARDNESS OF MINERALS<br />

Name Description or composition S<strong>pe</strong>cific gravity Mohr hardness<br />

Borax Na2B4O7·10H2O 1.7 2–2.5<br />

Bornite Cu5FeS4 5.0–5.1 3<br />

Boulangerite Pb5Sb4S11 6–6.3 2.5–3<br />

Boumonite PbCuSbS3 5.8–5.9 2.5–3<br />

Brannerite (U,Ca,Ce)(Ti,Fe) 2O6 4.5–5.4 4.5<br />

Braunite 3Mn2O3·MnSiO3 4.8 6–6.5<br />

Bravoite (Ni,Fe)S2 4.66 5.5–6<br />

Brochantite Cu4(OH) 6SO4 3.9 3.5–4<br />

Bromyrite Ag(Br,Cl)—Br,Cl 6–6.5 2.5<br />

Bronzite (Mg,Fe)SiO3 3.1–3.3 5.5<br />

Brookite TiO2 3.9–4.1 5.5–6<br />

Brucite Mg(OH) 2 2.39 2.5<br />

Bytownite Ab30An70-Ab10An90 2.74 6<br />

Caimgom Quartz, black/smoky<br />

Calamine Hemimorphite<br />

Calaverite AuTe2 9.35 2.5<br />

Calcite CaCO3 2.73 3<br />

Calomel Hg2Cl2 7.2 1.5<br />

Cancrinite (Fp) (Na2,Ca) 4(AlSiO4) 6CO3nH2O 2.45 5–6<br />

Carnalite KMgCl3·6H2O 1.6 1.0<br />

Carnolite K(UO2) 2(VO4) 2·3H2O 4.1 soft<br />

Cassiterite SnO2 6.8–7.1 6–7<br />

Cat’s eye Chrysoberyl or quartz, gemstone<br />

Celestite SrSO4 3.9–4.0 3–3.5<br />

Celsian (F) BaAl2Si2O8 3.37 6<br />

Cerargyrite Ag(Cl,Br)—Cl, Br 5.5–6 2.5<br />

Cerussite PbCO3 6.55 3–3.5<br />

Cervanite Sb2O4 4.0–5.0 4–5<br />

Chabazite (Z) Ca(Al2Si4O12)·6H2O 2.0–2.2 4–5<br />

Chalcanthite CuSO4·5H2O 2.1–2.3 2.5<br />

Chalcedony Cryptocrystaline quartz 2.6–2.7<br />

Chalcocite Cu2S 5.5–5.8 2.5–3<br />

Chalcopyrite Cu2FeS2 4.1–4.3 3.5–4<br />

Chalcotrichite Cuprite, fibrous<br />

Chalk Calcite, fine-grained<br />

Chalybite Siderite<br />

Chert SiO2, cryptocrystalline quartz 2.65<br />

Chessylite Azurite<br />

Chiastolite Andalusite<br />

Chloanthite Skutterudite, nickel variety<br />

Chlorite (MgFe2+ ,Fe3+ ) 6AlSi3O10(OH) 8 2.6–2.9 2–2.5<br />

Chloritoid (M) Fe2Al4Si2O10(OH) 4 3.5 6–7<br />

Chondrodite (Mg,Fe)3SiO4(OH,F) 2 3.1–3.2 6–6.5<br />

Chromite (Fe,Mg)O.(Fe,Al,Cr) 2O3 4.3–4.6 5.5<br />

Chrysoberyl BeAl2O4 3.6–3.8 8.5<br />

Chrysocolla Cu2H2(Si2O5)(OH) 4 2.0–2.4 2 -4<br />

Chrysolite Olivine<br />

Chrysoprase Chalcedony, green<br />

Chrysolite Ser<strong>pe</strong>ntine asbestos<br />

A.3


A.4 APPENDIX A<br />

Name Description or composition S<strong>pe</strong>cific gravity Mohr hardness<br />

Cinnabar HgS 8.10 2.5<br />

Cinnamon stone Grossularite garnet<br />

Citrine Quartz, pale yellow<br />

Clay A family of minerals<br />

Cleavelandite Albite, white<br />

Cliachite Aluminum hydroxide in bauxite<br />

Clinochlore Chlorite<br />

Clinoclase Cu3(AsO4)(OH) 3 4.38 2.5–3<br />

Clinoenstatite (E) (Mg,Fe)SiO3 3.19 6<br />

Clinoferrosilite (P) (Fe,Mg) SiO3 3.6 6<br />

Clinohumite Mg9Si4O16(F,OH) 2 3.1–3.2 6<br />

Clinozoisite Ca2Al3Si3O12(OH) 3.2–3.4 6–6.5<br />

Cobalite CoAsS 6.33 5.5<br />

Colemanite Ca2B6O11·5H2O 2.42 4–4.5<br />

Collophane Apatite<br />

Columbite (Fe,Mn)(Nb,Ta) 2O6-NbTa 5.2–6.7 6<br />

Cop<strong>pe</strong>r Cu 8.9 2.5–3<br />

Cop<strong>pe</strong>r glance Chalcocite<br />

Cop<strong>pe</strong>r pyrites Chalcopyrite<br />

Cordierite (Mg,Fe) 2Al4Si5O18 2.6–2.7 7–7.5<br />

Corundum Al2O3 4.0 9<br />

Covellite CuS 4.6–4.7 1.5–2<br />

Cristobalite SiO2, high-tem<strong>pe</strong>rature quartz 2.3 7<br />

Crocidolite Na3Fe2+ 3Fe3+ 2(SiO23)(OH) 3.2–3.3 4<br />

Crocoite PbCrO4 5.9–6.1 2.5–3<br />

Cryolite Na3AlF6 2.9–3 2.5<br />

Cubanite CuFe2S3 4.0–4.2 3.5<br />

Cummingtonite (Fe,Mg) 7(Si8O22)(OH) 2 3.1–3.6 6<br />

Cuprite Cu2O Cyanite Kyanite<br />

Cymophane Chrysoberyl<br />

Danaite (Fe,Co)As 5.9–6.2 5.5–6<br />

Danburite CaB2(SiO4) 2 2.9–3.0 7.0<br />

Datolite CaB(SiO4)(OH) 2.8–3.0 5.0–5.5<br />

Davidite Brannerite, Th variety<br />

Demantoid (G) Andradite garnet, green gemstone<br />

Diallage Diopside<br />

Diamond C 3.5 10<br />

Diaspore AlO(OH) 3.3–3.4 6.5–7.0<br />

Diatomite Diatoms, siliceous 0.4–0.6 2<br />

Dichroite Cordierite<br />

Dickite (C) Al2Si2O5(OH) 4, kaolin 2.6 2.0–2.5<br />

Digenite Cu9S5 5.6 2.5–3.0<br />

Diopside (P) CaMg(SiO3) 2 3.2–3.3 5.0–6.0<br />

Dioptase CuSiO2(OH) 2 3.3 5.0<br />

Disthene Kyanite<br />

Dolomite CaMg(CO3) 2 2.85 3.5–4<br />

Dry bone ore Smithsonite<br />

Dumortierite (Al,Fe) 7O3(BO3)(SiO4) 3 3.2–3.4 7.0


SPECIFIC GRAVITY AND HARDNESS OF MINERALS<br />

Name Description or composition S<strong>pe</strong>cific gravity Mohr hardness<br />

Endenite (H) Ca2NaMg5(AlSi7O22) 3.0 6.0<br />

Electrum Au, Ag, natural alloy 13.5–17 3.0<br />

Eleolite Nepheline<br />

Embolite Ag(Cl,Br)—Cl=Br 5.6 1.0–1.5<br />

Emerald Beryl, green gemstone<br />

Emery Corundum with magnetite<br />

Enargite Cu3AsS4 4.4–4.5 3.0<br />

Endichite Vanadite, arsenic variety<br />

Enstatite (P) MgSiO3 3.2–3.5 5.5<br />

Epidote Ca2(Al,Fe) 3Si3O12(OH) 3.3–3.5 6–7<br />

Epsomite MgSO4·7H2O, Epsom salt 1.75 2.0–2.5<br />

Erythrite Co3(AsO4)·8H2O 2.95 1.5–2.5<br />

Essonite (G) Grossularite<br />

Euclase BeAlSiO4(OH) 3.1 7.5<br />

Eucryptite (Y,Ce,Ca,U,Th) 2(Ti,Nb,Ta,Fe) 2O6 5.0–5.9 5.5–5.6<br />

Fahlore Tetrahedrite<br />

Fayalite (OV) Fe2SiO4 4.14 6.5<br />

Feather ore Jamesonite<br />

Feldspar (F) A group of minerals<br />

Feldspathoid A group of minerals<br />

Ferberite (W) FeWO4 7.5 5<br />

Fergusonite (R) (RE,Fe)(Nb,Ta,Ti) O4 4.2–5.8 5.5–6.5<br />

Ferrimolybdite Fe2(MoO4) 3·8H2O 3 1.5<br />

Ferrosilite (P) FeSiO3 3.6 6<br />

Fibrolite Sillimanite<br />

Flint SiO2, cryptocrystalline quartz 2.65 7<br />

Flos ferri Aragonite, arborescent<br />

Fluorite CaF2 Fool’s gold Pyrite<br />

Formanite (R) Fergusonite with TaNb<br />

Forsterite (Ov) Mg2SiO4 3.2 6.5<br />

Fowlerite Rhodonite, zinc bearing<br />

Franklinite (Fe2+ ,Zn,Mn2+ ) (Fe3+ , Mn3+) 2O4 5.15 6<br />

Freibergite Tetrahedrite, silver bearing<br />

Gadolinite (R) Be2FeY2Si2O10 4.0–4.5 6.5–7.0<br />

Gahnite (S) ZnAl2O4 4.55 7.5–8.0<br />

Galaxite (S) MnAl2O4 4.03 7.5–8.0<br />

Galena PbS 7.4–7.6 2.5<br />

Garnet (G) A group of minerals 3.5–4.3 6.5–7.5<br />

Gamierite (Ni,Mg) 3Si2O5(OH) 4 2.2–2.8 2.0–3.0<br />

Gaylussite Na2Ca(CO3) 2·5H2O 1.99 2.0–3.0<br />

Gedrite (A) Anthophylite, Al variety<br />

Geocronite Pb5(Sb,As) 2S8 6.6–6.5 2.5<br />

Gersdorffite NiAsS 5.9 5.5<br />

Geyserite Opal<br />

Gibbsite Al(OH) 3 2.3–2.4 2.5–3.5<br />

Glauberite Na2Ca(SO4) 2 2.7–2.8 2.5–3.0<br />

Glaucodot Danaite<br />

Glauconite (M) (K,Na)(Al,Fe3+ ,Mg) 2(Al,Si) 4O10(OH) 2 2.3 2<br />

A.5


A.6 APPENDIX A<br />

Name Description or composition S<strong>pe</strong>cific gravity Mohr hardness<br />

Glaucophane (A) Na2(Mg,Fe2+ ) 3Al2Si8O22(OH) 2 3.0–3.2 6.0–6.5<br />

Gmelinite (Z) (Na2·Ca)Al2Si4O12·6H2O 2.0–2.2 4.5<br />

Goethite FeO(OH) 4.37 5.0–5.5<br />

Gold Au 15–19.3 2.5–3.0<br />

Goslarite ZnSO4·7H2O 1.98 2.0–2.5<br />

Graphite C 2.3 1.0–2.0<br />

Greenockite CdS 4.9 3.0–3.5<br />

Grossularite (G) Ca3Al2(SiO4) 3 3.53 6.5<br />

Gummite UO3·nH2O 3.9–6.4 2.5–5<br />

Gypsum CaSO4·2H2O 2.32 2.0<br />

Halite NaCl, common salt 2.16 2.5<br />

Hallosite (C) Al2Si2O5(OH)·nH2O 2.0–2.2 1.0–2.0<br />

Harmotome (Z) (Ba,K)(Al,Si) 2Si6O16·6 H2O 2.45 4.5<br />

Hastingsite (H) NaCa2(Fe,Mg) 5Al2Si6O22(OH) 2 3.2 6.0<br />

Hausmannite Mn3O4 4.8 5.5<br />

Hauynite (Fp) (Na,Ca) 4·8Al6Si6O24·(SO4·S) 1-2 2.4–2.5 5.5–6<br />

Hectorite (C) (Mg,Li) 6Si8O20(OH) 4 2.5 1.0–1.5<br />

Hedenbergite (P) CaFe(Si2O6) 3.6 5.0–6.0<br />

Heliotro<strong>pe</strong> Chalcedony, green and red<br />

Helvite (Mn,Fe,Zn) 4Be3(SiO4) 3S 3.2–3.4 6.0–6.5<br />

Hematite Fe2O3 5.3 5.5–6.5<br />

Hemimorphite Zn4(Si2O7)(OH) 2·H2O 3.4–3.5 4.5–5.0<br />

Hercynite (S) FeAl2O4 4.4 7.5–8.0<br />

Hessite Ag2Te 8.4 2.5–3.0<br />

Heulandite (Na,Ca) 4—6Al6(Al,Si) 4Si26O72·24H2O 2.2 3.5–4.0<br />

Hiddenite Spodumene, green<br />

Holmquisite (A) Glaucophane, Li variety<br />

Homblende Ca2Na(MgFe2+ ) 4(AlFe3+ ,Ti)Al Si8O22 (O,OH) 2<br />

3.2 5.6<br />

Hom silver Cerargyrite<br />

Huebnerite (W) MnWO4 7.0 5.0<br />

Humite Mg7(SiO4) 3(F,OH) 2 3.1–3.2 6.0<br />

Hyacinth Zircon<br />

Hyalite Opal<br />

Hyalophane (O) (K,Ba)Al(Al,Si) 3O8 2.8 6.0<br />

Hydromica (M) Illite<br />

Hydrozincite Zn5(CO3) 2(OH) 6 3.6–3.8 2.0–2.5<br />

Hy<strong>pe</strong>rsthene (P) (MgFe)SiO3 3.4–3.5 5.0–6.0<br />

Ice H2O 0.95 1.5<br />

Iddingsite H8Mg9Fe2Si3O14 3.5–3.8 3.0<br />

Idocrase Ca10(Mg,Fe) 2Al4(SiO4) 5(Si2O7) 2(OH) 4 3.3–3.4 6.5<br />

Illite (C) Al,K,Ca,Mg<br />

Ilmenite FeTiO3 4.7 5.5–6.0<br />

Ilvaite CaFe2+ 2Fe3+ (SiO4) 2(OH) 4.0 5.5–6.0<br />

Indicolite Tourmaline<br />

Iodobromite Ag(Cl,Br,I) 5.7 1.0–1.5<br />

Iodyrite AgI 5.7 1.0–1.5<br />

Iolite Cordierite, gemstone<br />

Iridium Ir, platinoid 22.7 6.0–7.0


SPECIFIC GRAVITY AND HARDNESS OF MINERALS<br />

Name Description or composition S<strong>pe</strong>cific gravity Mohr hardness<br />

Iridosmine Ir,Os, platinoid 19.3–21.0 6.0–7.0<br />

Iron pyrite Pyrite<br />

Jacinth Hyacinth, zircon<br />

Jacobsite (S) (Mn2+ ,Fe2+ ,Mg)(Fe3+ ,Mn3+ ) 2O4 5.1 5.5–6.5<br />

Jade Jadeite or nephrite<br />

Jadeite (P) Na(Al,Fe)Si2O6 3.3–3.5 6.5–7.0<br />

Jamesonite Pb4FeSb6S14 5.5–6.0 2.0–3.0<br />

Jarcon Zircon<br />

Jarosite KFe3(SO4) 2(OH) 6 2.9–3.3 3.0<br />

Jas<strong>pe</strong>r Quartz<br />

Kainite MgSO4·KCl.3H2O 2.1 3.0<br />

Kalinite Alum, potash<br />

Kallophilite K(AlSiO4) 2.61 6.0<br />

Kalsilite Nephelines<br />

Kaolin Clay mineral<br />

Kaolinite Al2(Si2O5)(OH) 4 2.6–2.7 2.0–2.5<br />

Kemite Na2B4O7·4H2O 1.95 3.0<br />

Krennerite AuTe2 8.62 2.0–3.0<br />

Kunzite Spudomene, pink<br />

Kyanite Al2SiO5 3.6–3.7 5.0–7.0<br />

Labradorite (P) Ab50An50·Ab30An70 2.71 6.0<br />

Langbeinite K2Mg2(SO4) 3 2.83 2.5–3.5<br />

Lapis lazuli Impure lazurite<br />

Larsenite (Ov) PbZnSiO4 5.9 3.0<br />

Laumontite (Z) (Ca,Na)Al2Si4O12·4H2O 2.3 4.0<br />

Lawsonite CaAl2(Si2O7)(OH) 2·H2O 3.1 8.0<br />

Lazulite (Mg,Fe3+ )Al2PO4) 2(OH) 2 3.0–3.1 5.0–5.5<br />

Lazurite (Na,Ca) 4(AlSiO4) 3(SO4,S,Cl) 2.4–2.5 5.0–5.5<br />

Lechatelierite SiO2, fused silica 2.2 6.0–7.0<br />

Lepidocrocite FeO(OH) 4.1 5.0<br />

Lepidolite (M)<br />

Leucite<br />

K(LiAl) 3(Si,Al) 4O10(F,OH) 2 2.8–3.0 2.5–4.0<br />

Libethenite Cu2(PO4)(OH) 4.0 4.0<br />

Limonite FeO(OH)·nH2O 3.6–4.0 5.0–5.5<br />

Linarite PbCu(SO4)(OH) 2 5.3 2.5<br />

Linnaeite Co3S4 4.8 4.5–5.5<br />

Lithium mica Lepidolite<br />

Lithiophilite Li(Mn2+ ,Fe2+ )PO4 3.5 5.0<br />

Loellingite FeAs2 7.4–7.5 5.0–5.5<br />

Magnesite MgCO3 3.0–3.2 3.5–5.0<br />

Magnetite (S) (Fe,Mg)Fe2O4 5.2 6.0<br />

Malachite Cu2CO3(OH) 2 3.9–4.0 3.5–4.0<br />

Manganite MnO(OH) 4.3 4.0<br />

Manganosite MnO 5.0–5.4 5.5<br />

Marcasite FeS2, white iron pyrite 4.9 6.0–6.5<br />

Margarite (M) CaAl2(Al2Si2O10)(OH) 12 3.0–3.1 3.5–5.0<br />

Marialite (Sc) 3NaAlSi3O8·NaCl 2.7 5.5–6.0<br />

Marmatite Sphalerite, iron bearing 3.9–4.0<br />

Martite Hematite after magnetite<br />

A.7


A.8 APPENDIX A<br />

Name Description or composition S<strong>pe</strong>cific gravity Mohr hardness<br />

Meerschaum Sepiolite<br />

Meionite (Sc) 3CaAl2Si2O8·CaCO3 2.7 5.0–5.6<br />

Melaconite Tenorite<br />

Melanite (G) Andradite garnet (black)<br />

Melanterite FeSO4·7H2O 1.9 2.0<br />

Melilite (Na,Ca) 2(Mg,Al)(Si,Al) 2O7 2.9–3.1 5.0<br />

Menaccanite Limenite<br />

Menaghinite (J) CuPb13Sb7S24 6.36 2.5<br />

Mercury Hg 13.6<br />

Miargyrite AgSbS2 5.2–5.3 2.5<br />

Mica (M) A group of minerals<br />

Microcline (F) K(AlSi3O8), K feldspar 2.5–2.6 6.0<br />

Microlite (Na,Ca) 2(Ta,Nb) 2O6 6.33 5.5<br />

Micro<strong>pe</strong>rthite (F) Microcline and albite<br />

Milerite NiS 5.3 -5.7 3.0–3.5<br />

Mimetite Pb5Cl(AsO4) 3 7.0–7.2 3.5<br />

Minium Pb3O4 8.9–9.2 2.5<br />

Mispickel Arsenopyrite<br />

Molybdenite MoS2 4.6–4.7 1.0–1.5<br />

Monazite (Ce,La,Y,Th)(PO4,SiO4) 5.0–5.3 5.0–5.5<br />

Monticellite CaMgSiO4, rare olivine 3.2 5.0<br />

Montmorillonite (C) (Al,Mg) 8(Si4O10) 3(OH) 10·10H2O 2.5 1.0–1.5<br />

Moonstone (O) Opalescent albite or orthoclase<br />

Morganite Beryl, rose color<br />

Mullite Al6Si2O13 3.2 6.0–7.0<br />

Muscovite (M) KAl2(AlSi3O10)(OH) 2 2.7–3.1 2.0–2.5<br />

Nacrite (C) Al2(Si2O5)(OH) 2, kaolin group 2.6 2.0–2.5<br />

Nagyagite Pb5Au(Te,Sb) 4S5-8 7.4 1.0–1.5<br />

Natroalunite Alunite with NaK<br />

Natrolite Na2(Al2Si3O10)·2H2O 2.3 5.0–5.5<br />

Nepheline (Fp) (Na,K)AlSiO4 2.5–2.7 5.5–6.0<br />

Nephrite Tremolite, similar to jade<br />

Niccolite NiAs 7.8 5.0–5.5<br />

Nickel bloom Annabergite<br />

Nickel iron Ni,Fe, meteorite alloy 7.8–8.2 5.0<br />

Ni skutterudite (Ni,Co,Fe)As3 6.1–6.9 5.5–6.0<br />

Nitre KNO3, salt<strong>pe</strong>ter 2.0–2.1 2.0<br />

Nontronite (C) Fe(AlSi) 8O20(OH) 4 2.5 1.0–1.8<br />

Norbergite Mg3(SiO4)(F,OH) 2 3.1–3.2 6.0<br />

Noselite (Fp) Na8Al6Si6O24(SO4) 2.2–2.4 6.0<br />

Octahedrite Anatase<br />

Oligoclase (P) Ab90An10·Ab70An30 2.5 6.0<br />

Olivine (Ov) (Mg,Fe) 2SiO4 3.3–4.4 6.5–7.0<br />

Onyx Chalcedony, layered structure<br />

Opal SiO2·nH2O 1.9–2.2 5.0–6.0<br />

Orpiment As2S3 3.5 1.5–2.0<br />

Orthite Allanite<br />

Orthoclase (F) K(AlSi3O8), K feldspar 2.6 6.0<br />

Osmiridium Iridosmine


SPECIFIC GRAVITY AND HARDNESS OF MINERALS<br />

Name Description or composition S<strong>pe</strong>cific gravity Mohr hardness<br />

Ottrelite (M) (Fe2+ ,Mn)(Al,Fe3+ )Si3O10·H2O 3.5 6.0–7.0<br />

Palladium Pd 11.9 4.5–5.0<br />

Paragonite (M) NaAl2(AlSi3O10)(OH) 12 2.85 2.0<br />

Pargasite (H) NaCa2Mg4Al3Si6O22(OH) 2 3.0–3.5 5.5<br />

Peacock ore Bomite<br />

Pearceite (Ag,Cu) 16As2S11 6.15 3.0<br />

Pectolite NaCa2Si3O8(OH) 2.7–2.8 5.0<br />

Penninite Chlorite<br />

Pentlandite (Fe,Ni) 9S8 4.6–5.0 3.5–4.0<br />

Peridot (Ov) Olivine, gemstone<br />

Perovskite CaTiO3 4.03 5.5<br />

Perthite (F) Mixture of microcline and albite<br />

Petalite (Fp) Li(AlSi4O10) 2.4 6.0–6.5<br />

Petzite Ag3AuTe2 8.7–9.0 2.5–3.0<br />

Phenacite Be2SiO4 2.9–3.0 7.5–8.0<br />

Phillipsite (Z) (K2,Na2,Ca)Al2Si4O12·H2O 2.2 4.5–5.0<br />

Phlogopite (M) K(Mg,Fe) 3AlSi3O10(OH,F) 2 2.86 2.5–3.0<br />

Phosgenite Pb2Cl2CO3 6.0–6.3 3.0<br />

Phosphuranylite Ca(UO2) 4(PO4) 2(OH) 4·7H2O Picotite (S) Spinel, chromium<br />

Piedmontite Epidote, Mn2+ 3.4 6.5<br />

Pigeonite (P) (Ca,Mg,Fe)SiO3 3.2–3.4 5.0–6.0<br />

Pinite (M) Muscovite mica<br />

Pitchblende Uraninite<br />

Plagioclase (P) A group of aluminum silicates<br />

Plagionite (J) Pb5Sb8S17 5.6 2.5<br />

Platinum Platinum metal alloy 14–19 4.0–4.5<br />

Pleonaste (S) Spinel, iron<br />

Plumbago Graphite<br />

Polianite MnO2, pyrolusite 5.0 6.0–6.5<br />

Pollucite (Cs,Na) 2Al2Si4O12·H2O 2.9 6.5<br />

Polybasite (Ag,Cu) 16Sb2S11 6.0–6.2 2.0–3.0<br />

Polycrase (R) Y,Ce,Ca,U,Th,Ti,Nb,Ta,Fe oxide 4.7–5.9 5.5–6.5<br />

Polyhalite K2Ca2Mg(SO4) 4·2H2O 2.78 2.5–3.0<br />

Potash aluminum KAl(SiO4) 2·11H2O 1.75 2.0–2.5<br />

Potassium feld KalSi3O8, see orthoclase<br />

Potash mica (M) Muscovite<br />

Powellite CaMoO4 4.2 3.5–4.0<br />

Prase Jas<strong>pe</strong>r, green<br />

Prehnite Ca2Al2(Si3O10)(OH) 12 2.8–2.9 6.0–6.5<br />

Prochlorite Chlorite group<br />

Proustite Ag3AsS3 5.6 2.0–2.5<br />

Psilomelane Manganese mineral group<br />

Pyrargyrite Ag3SbS3 5.9 2.5<br />

Pyrite FeS2 5.02 6.0–6.5<br />

Pyrochlore (Na,Ca) 2(Nb,Ta) 2O6(OH,F) 4.2–4.5 5.0<br />

Pyrolusite MnO2 4.8 1–2<br />

Pyromorphite Pb5(PO4) 3Cl 6.5–7.1 3.5–4.0<br />

Pyro<strong>pe</strong> (G) (Mg,Fe) 3Al2(SiO4) 3 3.5 7.0<br />

A.9


A.10 APPENDIX A<br />

Name Description or composition S<strong>pe</strong>cific gravity Mohr hardness<br />

Pyrophylite Al2Si4O10(OH) 2 2.8–2.9 1.0–2.0<br />

Pyroxene (P) A group of minerals<br />

Pyrrhotite Fe1-x S where x = 0.0 to 0.2 4.6 4.0<br />

Quartz SiO2 2.7 7<br />

Rammelsbergite NiAs2 7.1 5.5–6.0<br />

Rasorite Kemite<br />

Realgar AsS 3.5 1.5–2.0<br />

Red ochre Hematite<br />

Rhodochrosite MnCO3 3.5–3.6 3.5–4.5<br />

Rhodolite (G) 3(Mg,Fe)O,Al2O3·3SiO2 3.8 7.0<br />

Rhodonite MnSiO3 3.6–3.7 5.5–6.0<br />

Riebeckite (A) Na2(Fe,Mg) 5Si8O22(OH) 2 3.4 4.0<br />

Rock salt Halite<br />

Roscoelite (M) K(U,Al,Mg) 3Si3O10(OH) 2 3.0 2.5<br />

Rubellite Tourmaline, red or pink<br />

Ruby Corundum, red, gemstone<br />

Ruby cop<strong>pe</strong>r Cuprite<br />

Ruby silver Pyrargyrite or proustite<br />

Rutile TiO2 4.2–4.3 6.0–6.5<br />

Samarskite (Y,Ce,U,Ca,Fe,Pb,Th)(Nb,Ta,Ti,Sn) 2O6 4.1–6.2 5.0–6.0<br />

Sanadine (O) Orthoclase, high tem<strong>pe</strong>rature<br />

Saponite (C) (Mg,Al) 6(Si,Al) 8O20(OH) 4 2.5 1.0–1.5<br />

Sapphire Corundum, blue, gemstone<br />

Satin spar Gypsum, fibrous<br />

Scapolite (Na or Ca) 4Al3(Al,Si) 3Si6O24(Cl,CO3,SO4) 2.6–2.7 5.0–6.0<br />

Scheelite CaWO4 5.9–6.1 4.5–5.0<br />

Schorlite Tourmaline, black<br />

Scolecite (Z) Ca(Al2Si3O10)·3H2O 2.2–2.4 5.0–5.5<br />

Scorodite FeAsO4·2H2O 3.1–3.3 3.5–4.0<br />

Scorzalite (Fe,Mg)Al2(PO4) 2(OH) 2 3.4 5.5–6.0<br />

Selenite Clear and crystalline gypsum<br />

Semseyite Pb9Sb8S21 5.8 2.5<br />

Sepiolite Mg4(Si2O5) 3(OH) 2·6H2O, Meerschaum 2.0 2.0–2.5<br />

Sericite (M) Muscovite mica, fine grained<br />

Ser<strong>pe</strong>ntine (Mg,Fe) 3SiO5(OH) 4 2.2 2.0–5.0<br />

Siderite FeCO3 3.8–3.9 3.5–4.0<br />

Siegenite (Co,Ni) 3S4 4.8 4.5–5.5<br />

Sillimanite Al2SiO5 3.2 6.0–7.0<br />

Silver Ag 10.5 2.5–3.0<br />

Silver glance Argentite<br />

Sklodowskite Mg(UO2) 2Si2O7·6H2O 3.5<br />

Skutterudite (Co,Ni,Fe)As3 6.1–6.9 5.0<br />

Smaltite Skutterudite variety<br />

Smithsonite ZnCO3 4.3–4.4 5.0<br />

Soapstone Talc<br />

Sodalite (Fp) Na4Al3Si3O12Cl 2.2–2.3 5.5–6.0<br />

Soda nitre NaNO3 2.3 1.0–2.0<br />

S<strong>pe</strong>cular iron Hematite, foliated<br />

S<strong>pe</strong>rrylite PtAs2 10.5 6.0–7.0


SPECIFIC GRAVITY AND HARDNESS OF MINERALS<br />

A.11<br />

Name Description or composition S<strong>pe</strong>cific gravity Mohr hardness<br />

S<strong>pe</strong>ssartite (G) Mn3Al2(SiO4) 3, red, brown 4.2 7.0<br />

Sphalerite (Zn,Fe)S 3.9–4.1 3.5–4.0<br />

Sphene CaTiO(SiO4) 3.4–3.5 5.0–5.5<br />

Spinel group (Mg,Fe,Zn,Mn)Al2O4 3.6–4.0 8.0<br />

Spodumene (P) LiAl(Si2O6) 3.1–3.2 6.5–7.0<br />

Stannite Cu2FeSnS4 4.4 4.0<br />

Staurolite (Fe,Mg) 2Al2Si4O23(OH) 3.6–3.8 7.0–7.5<br />

Steatite Talc<br />

Stephanite Ag5SbS4 6.2–6.3 2.0–2.5<br />

Stembergite AgFe2S3 4.1–4.2 1.0–1.5<br />

Stibnite Sb2S3 4.5–4.6 2.0<br />

Stilbite (Z) NaCa2Al5Si3O36·14H2O 2.1–2.2 3.5–4.0<br />

Stillwellite (Ce,La,Ba)BSiO5 4.6<br />

Stolzite PbWO4 8.3–8.4 2.5–3.0<br />

Stromeyerite (Cu,Ag)S 6.2–6.3 2.5–3.0<br />

Strontianite SrCO3 3.7 3.5–4.0<br />

Sulphur S 2.0–2.1 1.5–2.5<br />

Sunstone (F) Oliglase, translucent<br />

Sylvanite (Au,Ag)Te2 8.0–8.2 1.5–2.0<br />

Sylvite KCl 2.0 2.0<br />

Talc Mg3(Si4O10)(OH) 2 2.7–2.8 1.0<br />

Tantalite (Fe,Mn)(Ta,Nb) 2O6,TaNb 6.2–8.0 6.0–6.5<br />

Tennantite (Cu,Fe,Zn,Ag) 12As4S13 4.6–5.1 3.0–4.5<br />

Tenorite CuO 5.8–6.4 3.0–4.0<br />

Tephroite (Ov) Mn2(SiO4) 4.1 6.0<br />

Tetrahedrite (Cu,Fe,Zn,Ag) 12Sb4S13 4.6–5.1 3.0–4.5<br />

Thenardite Na2SO4 2.68 2.5<br />

Thomsonite (Z) NaCa2Al5Si5O20·6H2O 2.3 5.0<br />

Thorianite ThO2 9.7 6.5<br />

Thorite Th(SiO4) 5.3 5.0<br />

Thulite Zoisite, pink to red<br />

Tiger’s eye Form of quartz<br />

Tin Sn 7.3 2.0<br />

Tinstone Cassiterite<br />

Titanite Sphene<br />

Topaz Al2(SiO4)(F,OH) 2 3.4–3.6 8.0<br />

Torbernite Cu(UO2) 2(PO4) 2·8H2O 3.2 7.0–7.5<br />

Tourmaline (Na,Ca)(Al,Fe,Li,Mg) 3Al6(BO3) 3(Si6O22) (OH) 4<br />

3.0–3.2 7.0–7.5<br />

Tremolite (A) Ca2Mg5(Si8O22)(OH) 2 3.0–3.3 5.0–6.0<br />

Tridymite SiO2 2.3 7.0<br />

Triphylite Li(Fe,Mn)PO4 3.4–3.6 4.0–5.5<br />

Troilite Pyrrhotite<br />

Trona Na2CO3·NaHCO3·2H2O 2.1 3.0<br />

Troostite Manganiferous willemite<br />

Tungstite WO3·nH2O 2.5<br />

Turgite 2Fe2O3·nH2O 4.2–4.6 6.5<br />

Turquoise CuAl6(PO4) 4(OH) 8·5H2O 2.6–2.8 6.0<br />

Tyuyamunite Ca(UO2) 2(VO4) 2·5H2O 3.7–4.3 2.0


A.12 APPENDIX A<br />

Name Description or composition S<strong>pe</strong>cific gravity Mohr hardness<br />

Ulexite NaCaB5O9·8H2O 1.96 1.0<br />

Uralite (H) Homblende after pyroxene<br />

Uraninite UO2 to UO3 9.0–9.7 5.5<br />

Uranophane Ca(UO2) 2Si2O7·6H2O 3.8–3.9 2.0–3.0<br />

Uvarovite (G) Ca3Cr2(SiO4) 2, green 3.5 7.5<br />

Vanadinite Pb5(VO4) 3Cl 6.7–7.1 3.0<br />

Variscite Al(PO4)·2H2O 2.4–2.6 3.5–4.5<br />

Vermiculite (M) Biotite, altered 2.4 1.5<br />

Vesuvianite Idocrase<br />

Violarite Ni2FeS4 4.8 4.5–5.5<br />

Vivianite Fe3(PO4) 2·8H2O 2.6–2.7 1.5–2.0<br />

Wad Manganese oxides<br />

Wavelite Al3(OH) 3(PO4) 2·5H2O 2.3 3.5–4.5<br />

Wemerite (Sc) Scapolite<br />

White pyrite Marcasite<br />

White mica (M) Muscovite<br />

Wilemite Zn2SiO4 3.9–4.2 5.5<br />

Witherite BaCO3 4.3 3.5<br />

Wolframite (Fe,Mn)WO4 7.0–7.5 5.0–5.5<br />

Wollastonite Ca(SiO3) 2.8–2.9 5.0–5.5<br />

Wood tin Cassiterite<br />

Wulfenite PbMoO4 6.5–7.5 3.0<br />

Wurtzite (Zn,Fe)S 4.0 4.0<br />

Xenotime YPO4 4.4–5.1 4.0–5.0<br />

Zeolite A group of minerals<br />

Zincite ZnO 5.7 4.0–4.5<br />

Zinc spinel Gahnite<br />

Zinkenite (J) Pb6Sb14S27 5.3 3.0–3.5<br />

Zinnwaldite Fe,Li, mica 3.0 2.5–3.0<br />

Zircon ZrSiO4 4.7 7.5<br />

Zoisite Ca2Al3Si3O12(OH) 3.3 6.0<br />

Adapted from T. J. Glover, Pocket Ref, Second Edition. Littleton, CO: Sequoia.


CHAPTER 10<br />

MATERIALS SCIENCE FOR<br />

SLURRY SYSTEMS<br />

10-0 INTRODUCTION<br />

Slurry is essentially a mixture of solids and liquids. Abrasion, erosion, and corrosion are<br />

so associated with pumping <strong>slurry</strong> that manufacturers of <strong>slurry</strong> pumps sometimes s<strong>pe</strong>nd<br />

more money dealing with wear issues than with developing new hydraulics.<br />

Wear is very complex and too often oversimplified. It de<strong>pe</strong>nds on many factors such as<br />

the microstructure of the surface of the pump part, the hardness and sha<strong>pe</strong> of the crushed<br />

or milled minerals, the s<strong>pe</strong>ed of flow, the scaling of pi<strong>pe</strong>s, etc. Dredge and <strong>slurry</strong> pumps,<br />

ball mill liners and shells, magnetic separators, and agitators are made from metals that<br />

combine the ability to resist stress and impact loads, erosion, and corrosion. Excellent<br />

books on materials science are available to the reader but, unfortunately, they too often<br />

dedicate just few lines or one or two paragraphs to the white irons or polymers used by<br />

the designers of <strong>slurry</strong> <strong>systems</strong>. These materials are too often classified as materials for<br />

s<strong>pe</strong>cial applications. This chapter will therefore make an effort to expand on this topic, as<br />

a good understanding of it can save on maintenance costs to the o<strong>pe</strong>rator and is necessary<br />

for the successful design of a <strong>slurry</strong> system.<br />

10-1 THE STRESS–STRAIN RELATIONSHIP<br />

OF METALS<br />

The stress–strain relationship of metals under tension (Figure 10-1) is often represented in<br />

the form of a graph of stress versus strain. Stress is essentially the load force <strong>pe</strong>r unit area.<br />

It is called direct stress � when the load is normal to the force, and shear stress � when the<br />

load is parallel to the surface:<br />

� = FN/A (10-1)<br />

where<br />

� = direct stress<br />

� = tangential force<br />

A = area<br />

� = F N/A (10-2)<br />

10.1


10.2 CHAPTER TEN<br />

Stress<br />

u<br />

Yield y<br />

Elastic<br />

Limit<br />

e<br />

e<br />

E<br />

Utimate Tensile Strength<br />

2% offset<br />

strain<br />

FIGURE 10-1 Stress–strain relationship of metals.<br />

fracture<br />

FN = normal force<br />

FT = tangential force<br />

When a s<strong>pe</strong>cimen bar of steel is subject to a tension load at its ends (Figure 10-2), it<br />

stretches. Within a certain range, the elongation is elastic, and this means that if the load<br />

is removed, the s<strong>pe</strong>cimen will return to its original length. The maximum stress in this<br />

elastic range is called the elastic limit and is shown in Fig 10-1 as �e. The elongation �L<br />

is divided by the original length L0 of the s<strong>pe</strong>cimen to define the strain �:<br />

� = �L/L0 (10-3)<br />

It is a nondimensional measure, often expressed in <strong>pe</strong>rcentage of original length, and mistakenly<br />

called “elongation” instead of direct strain. The direct strain under tension is correlated<br />

to the direct stress in the elastic range up to �e by the Young modulus E:<br />

� = �/E (10-4)<br />

Steels, but not all metals, exhibit a further nonlinear elongation up to a value called the<br />

yield stress. The elongation becomes <strong>pe</strong>rmanent as the materials yield. Beyond the yield<br />

point, the elongation continues to grow, until a value called the ultimate tensile strength<br />

�u is reached. This is the point at which the s<strong>pe</strong>cimen can withstand the greatest load and<br />

beyond which fracture is likely to occur rather quickly.<br />

Direct stress by itself induces a degree of secondary shear stresses. For each material<br />

there is a Poisson ratio �. For steel it is 0.30 and for gray cast iron it is 0.26. Shear strain<br />

correlates with shear strain by the modulus of rigidity G (also called shear modulus). It is<br />

related to the Young modulus by the Poisson ratio:<br />

E<br />

G = � (10-5)<br />

2(1 + �)<br />

Due to fatigue considerations, steel shafts of rotating equipment are designed for maximum<br />

stresses of less than 18% to the ultimate strength or 30% of yield strength. This is<br />

due to the combination of direct stress, torsion, and bending moments. Components of a<br />

<strong>slurry</strong> mill or pump must be able to absorb the energy of impact without fracture. The energy<br />

due to impact involves both loads and deflections. The capacity to absorb such ener-<br />

f


gy is called the resilience. Its measure is the modulus of resilience. For steels, it is the area<br />

under the stress-to-strain triangle in the elastic range.<br />

10-2 IRON AND ITS ALLOYS FOR THE<br />

SLURRY INDUSTRY<br />

Cast iron is an alloy of iron, carbon, silicon, and manganese. Carbon is in the range of 2 to<br />

4%. The cooling rate after casting of cast iron and subsequent heat treatment determine its<br />

mechanical pro<strong>pe</strong>rties. Carbon is very important to the pro<strong>pe</strong>rties of cast iron.<br />

10-2-1 Grey Iron<br />

MATERIALS SCIENCE FOR SLURRY SYSTEMS<br />

10.3<br />

FIGURE 10-2 Hardness of minerals. (From Wilson, 1985. Reproduced by <strong>pe</strong>rmission from<br />

McGraw-Hill.)<br />

Grey iron is cast iron with carbon precipitated in the form of graphite flakes. Graphite<br />

flakes weaken the casting in tension, and grey iron is considered to have a compressive


10.4 CHAPTER TEN<br />

strength three to five times as much as in tension. Grey iron has good damping pro<strong>pe</strong>rties<br />

and is used in the base of machinery, bearing assembly of <strong>slurry</strong> pumps, parts essentially<br />

under compression such as certain low-pressure pump casings, engine blocks, gears, flywheels,<br />

and brake disks.<br />

10-2-2 Ductile Iron<br />

The addition of magnesium as an alloying element precipitates excess carbon in the form<br />

of small nodules. These nodules do not disturb the structure of cast iron, as is the case<br />

with the graphite flakes. Ductile iron, also called nodular iron, has better pro<strong>pe</strong>rties in tension,<br />

and has better ductility, impact resistance, and stiffness than grey iron. Ductile iron<br />

is used for the casting of high-pressure pump casings or metal- and rubber-lined pumps,<br />

as well as unlined casings for mild slurries. Ductile iron is available in different grades<br />

such as 60-40-18 or 60,000 psi (413 MPa) ultimate tensile strength, 40,000 psi (276 MPa)<br />

yield, and 18% elongation. Such an elongation is not required for <strong>slurry</strong> pumps, and following<br />

heat treatment and alloying, other grades are used. For pumps, ductile iron that<br />

meets the standard ASTM A536-84 or SAE-J434C is available with a tensile strength of<br />

552 MPa (80,000 psi), a yield strength of 414 MPa (60,000 psi), elongation of 3%, and<br />

with a minimum hardness of 187 BHN.<br />

10-3 WHITE IRON<br />

Wear-resistant alloy irons are essentially “white irons.” They have found widespread application<br />

in the mining industry for the manufacturing of crushers, mill liners, and <strong>slurry</strong><br />

pumps as well as for shot blasting. White irons used to be considered a very useless byproduct<br />

of the melting of irons, with all the carbon precipitating in the form of carbides in<br />

the <strong>pe</strong>arlitic matrix. Up to the beginning of the 20th century, they were discarded because<br />

they are extremely brittle and impossible to machine.<br />

10-3-1 Malleable Iron<br />

Malleable iron is made from white iron by a two-stage heat treatment process. The resultant<br />

structure contains excess graphite in the form of tem<strong>pe</strong>red nodules. Because white<br />

iron is used, castings can be thinner than 76 mm (3�). Malleable iron has found applications<br />

for the bearing surfaces of heavy parts of farm equipment, trucks, railroad equipment,<br />

and to a certain extent in some <strong>slurry</strong> applications.<br />

Malleable cast iron has a structure that consists of ferrite, <strong>pe</strong>arlite, and graphite. Its ultimate<br />

tensile strength is in the range of 400 to 500 MPa (58,000–72,500 psi). Its ductility<br />

and toughness decrease as the quantity of <strong>pe</strong>arlite increases. Zakharov (1962) described<br />

the conversion of white iron to malleable cast iron by a two-step heat treatment process.<br />

To be pro<strong>pe</strong>rly heat-treated, the carbon content must be low and must not exceed 2.5<br />

to 2.8%. The lower the carbon content, the less graphite forms. Silicon must not exceed<br />

1% and manganese must not exceed 0.5%. If the silicon content exceeds 1%, it prevents<br />

the transformation of graphite flakes into nodules. Although the presence of manganese<br />

facilitates the casting of white iron, excessive amounts tend to stabilize the carbides during<br />

heat treatment. Stabilized carbides increase the resistance to wear but they also make<br />

it very difficult to machine the cast component.


In the first step of heat treatment, the white iron is heated to 900°C–950°C<br />

(1650–1742°F) and then allowed to cool. During the second step, it is annealed at 720°C<br />

to 760°C (1330°F–1400°F). Before annealing, white iron consists essentially of two phases:<br />

austenite and cementite (a constituent of lebedurite eutectic). During annealing,<br />

austenite is not affected, but the cementite decomposes to form iron and graphite:<br />

Fe3C = 3Fe + C (graphite)<br />

After this first step of graphitization, the malleable cast iron consists of austenite and<br />

graphite. The carbon content of austenite is about 1% at 900°C (1650°F). If the white iron<br />

is allowed to cool after this first step of heat treatment, secondary cementite or ferrite<br />

forms, de<strong>pe</strong>nding on the applied cooling rate. If cementite and ferrite are considered to be<br />

undesirable from a point of view of machining, a second step of heat treatment is applied<br />

to decompose the secondary and <strong>pe</strong>arlite cementite and to form nodular carbon.<br />

Conventionally annealing white iron into malleable form used to be a very lengthy<br />

process that could take three days or more in a heat treatment furnace. Various methods<br />

have been develo<strong>pe</strong>d over the years to accelerate the process, such as first-step heat treatment<br />

at 1000°C (1832°F), hardening before annealing, etc.<br />

10-3-2 Low-Alloy White Irons<br />

The British Standard BS 4844:1986 defines three grades of low-alloy white irons shown<br />

in Tables 10-1 and 10-2. These alloys have been su<strong>pe</strong>rseded by alloyed irons.<br />

10-3-3 Ni-Hard<br />

MATERIALS SCIENCE FOR SLURRY SYSTEMS<br />

10.5<br />

The International Nickel Company develo<strong>pe</strong>d s<strong>pe</strong>cial alloys of white iron with nickel.<br />

These are called Ni-hard and a number of alloys such as Ni-hard 1 to Ni-hard 4 are now<br />

produced (Tables 10-3 to 10-6). The presence of nickel increases the hardness but it also<br />

ensures the transformation of the austenite to martensite after pro<strong>pe</strong>r heat treatment. The<br />

selection of alloying elements is based on the intended use and on the thickness of the cast<br />

part. The maximum carbon content is 3.2–3.6%, but when impact resistance is important,<br />

the carbon should be trimmed to 2.7–3.2%.<br />

The composition of Ni-hard 1 to 4 is not exactly the same from one country to another,<br />

as shown in table 10-3 to 10-6.<br />

Ni-hard 1, or ASTM A532 Class 1, Ty<strong>pe</strong> A, is a martensitic white iron. It is used in<br />

relatively mild erosive applications were impact forces are low. It is heat treated for stress<br />

relief. Due to the limitations on thickness to 200 mm (8�), as indicated in Table 10-3, Nihard<br />

1 has found limited applications in wastewater plants and mild slurries.<br />

Ni-hard 4 (ASTM A532 Class I, Ty<strong>pe</strong> D) has a tensile strength in the range of<br />

420–700 MPa (60,000 to 100,000 psi). To increase its hardness, some manufacturers of<br />

<strong>slurry</strong> pumps conduct cryogenic or heat treatment. Its excellent fluidity makes it a suitable<br />

TABLE 10-1 Composition of Low-Alloy White Irons<br />

BS 4844:1986 C Si Mn Cr<br />

Grade 1A, 1B, 1C 2.4–3.4 0.5–1.5 0.2–0.8 2.0 max


10.6 CHAPTER TEN<br />

TABLE 10-2 Hardness of Low-Alloy White Irons<br />

alloy for casting fairly complex sha<strong>pe</strong>s of wear-resistant liners for mills, grinding balls,<br />

and pump parts.<br />

10-3-4 High-Chrome–Molybdenum Alloys<br />

By adding chrome in the range of 12 to 28%, together with nickel and molybdenum, allows<br />

the casting of abrasion-resistant alloys that are tough and can be cast in large sizes to<br />

match the needs of the mining industry. Eutectic carbides in the form of M 7C 3 in combination<br />

with an austenitic, martensitic, or <strong>pe</strong>arlitic matrix gives a full range of alloys. Some<br />

of the components are cast <strong>pe</strong>arlitic to allow machining, then heat treated to obtain an<br />

abrasion-resistant martensitic structure. Tables 10-7 and 10-8 compare two alloys in this<br />

family as cast in France, Germany, the United Kingdom, and the United States. They are<br />

often called by the foundries 16% and 27% chrome.<br />

High-chrome white irons cannot be welded. They are often very difficult to machine<br />

TABLE 10-3 Composition of Ni-Hard 1 White Iron<br />

France Germany United Kingdom United States<br />

Standard NF 32-401 (1980) DIN 1695 (1981) BS 4844 Pt2:1972 ASTM A532<br />

(1982)<br />

Grade FB Ni4 Cr2 HC G-X330 NiCr4 2 2B Class 1 Ty<strong>pe</strong> A<br />

Ni-Cr LC<br />

C (%) 3.2–3.6 3.0–3.6 3.2–3.6 3.0–3.6<br />

Si (%) 0.2– 0.8 0.2–0.8 0.3–0.8 0.8 max<br />

Mn (%) 0.3–0.7 0.3–0.7 0.2–0.8 1.3 max<br />

Ni (%) 3.0–5.5 (may be<br />

replaced by Cu)<br />

3.3–5.0 3.0–5.5 3.0–5.0<br />

Cr (%) 1.5–2.5 1.4–2.4 1.5–2.5 1.4–4.0<br />

Cu (%) — — — —<br />

Mo (%) 0.0–1.0 0.5 0.5 max —<br />

P max (%) — 0.3 —<br />

S max (%) — 0.15 0.15<br />

Minimum hardness<br />

as cast<br />

— 450 (430) HB 550 (500) HB 550 HB<br />

Minimum hardness<br />

as cast<br />

500–700 HB 550 (500) 550<br />

Annealed —<br />

Maximum section 200 mm (8�)<br />

Data from Brown (1994).<br />

Hardness (minimum) HB<br />

Grade Thickness � 50 mm Thickness > 50 mm<br />

1A 400 350<br />

1B 400 350<br />

1C 250 200


MATERIALS SCIENCE FOR SLURRY SYSTEMS<br />

TABLE 10-4 Composition of Ni-Hard 2 White Iron<br />

10.7<br />

France Germany United Kingdom United States<br />

Standard NF 32-401 (1980) DIN 1695 (1981) BS 4844 Pt2:1972 ASTM A532<br />

(1982)<br />

Grade FB Ni4 Cr2 BC G-X260 NiCr4 2 2A Class 1 Ty<strong>pe</strong> B<br />

Ni-Cr LC<br />

C (%) 2.7–3.2 2.6–2.9 2.7–3.2 2.5–3.0<br />

Si (%) 0.2– 0.8 0.2–0.8 0.3–0.8 1.3 max<br />

Mn (%) 0.3–0.7 0.3–0.7 0.2–0.8 0.8 max<br />

Ni (%) 3.0–5.5 (may be<br />

replaced by Cu)<br />

3.3–5.0 3.0–5.5 3.0–5.0<br />

Cr (%) 1.5–2.5 1.4–2.4 1.5–2.5 1.4–4.0<br />

Cu (%) — — — —<br />

Mo (%) 0.0–1.0 0.5 max 1.0 max<br />

P max (%) — 0.3 0.3<br />

S max (%) — 0.15 0.15<br />

Minimum hardness<br />

as cast<br />

— 450 (430) HB 550 HB<br />

Minimum hardness<br />

as cast<br />

450–650 HB 520 (480) 500<br />

Annealed —<br />

Maximum section 200 mm (8�)<br />

Data from Brown (1994).<br />

except by grinding. An electric induction furnace is used to achieve the high tem<strong>pe</strong>rature<br />

needed for melting. The Cr–Fe oxides that form during the melting process attack the silica<br />

lining of furnaces. Alumina linings should be used for these furnaces.<br />

Scrap steel, foundry return, s<strong>pe</strong>nt im<strong>pe</strong>ller, and SAG mill liners are melted with addition<br />

of high- and low-carbon ferrochromium to obtain the iron at the required carbon con-<br />

TABLE 10-5 Composition of Ni-Hard 3 White Iron<br />

France United States<br />

Standard NF 32-401 (1980) ASTM A532 (1982)<br />

Grade FB A Class 1 Ty<strong>pe</strong> C Ni-Cr-GB<br />

C (%) 2.7–3.9 2.9–3.7<br />

Si (%) 0.4–1.5 1.3 max<br />

Mn (%) 0.2–0.8 0.8 max<br />

Ni (%) 0.3–3.0 2.7–4.0<br />

Cr (%) 0.2–2.0 1.1–1.5<br />

Cu (%) — —<br />

Mo (%) 0.0–1.0 1.0<br />

P max (%) — 0.3<br />

S max (%) — 0.15<br />

Minimum hardness as cast — 550 HB<br />

Minimum hardness as cast<br />

Annealed<br />

500–700 HB<br />

Maximum section 75 mm (3�) diameter ball<br />

Data from Brown (1994).


10.8 CHAPTER TEN<br />

TABLE 10-6 Composition of Ni-Hard 4 White Iron<br />

France Germany United Kingdom United States<br />

Standard NF 32-401<br />

(1980)<br />

DIN 1695<br />

(1981)<br />

BS 4844 Pt2:1972<br />

ASTM A532<br />

(1982)<br />

Grade FB Cr9 Ni5 GX 300 CrNSi 2C 2D 2E Class 1 Ty<strong>pe</strong> D<br />

9 5 2 Ni-Hi-Cr<br />

C (%) 2.5–3.6 2.5–3.5 2.4–2.8 2.8–3.2 3.2–3.6 2.5–3.6<br />

Si (%) 1.5–2.2 1.5–2.2 1.5–2.2 1.5–2.2 1.5–2.2 1.3 max<br />

Mn (%) 0.3– 0.7 0.3– 0.7 0.2– 0.8 0.2– 0.8 0.2– 0.8 1.0–2.2<br />

Ni (%) 4.0–6.0 8.0–10.0 4.0–6.0 4.0–6.0 4.0–6.0 4.5–7.0<br />

Cr (%) 5.0–11.0 4.5–6.5 8.0–10.0 8.0–10.0 8.0–10.0 7.0–10.0<br />

Cu (%) — —<br />

Mo (%) 0.5 max 0.5 0.5 max 0.5 max 0.5 max 1.0 max<br />

P max (%) — 0.3 0.3 0.3 0.3 0.1<br />

S max (%) — 0.15 0.15 0.15 0.15 0.15<br />

Minimum<br />

hardness<br />

as cast<br />

450 (430) HB 550 HB<br />

Minimum<br />

hardness<br />

as cast<br />

550–750 HB 600 (535) HB 500 550 600 600<br />

Annealed 400<br />

Maximum 300 mm (12�)<br />

section diameter ball<br />

Data from Brown (1994).<br />

tent. Ferromolybdenum is added. About 5% of the chrome is lost during melting. The<br />

melting tem<strong>pe</strong>rature for the low-carbon, high-chrome iron is in the range of 1600 to<br />

1650°C (or 2912 to 3002°F), but for the higher carbons it is in the range of 1550 to<br />

1600°C (or 2822 to 2912°F).<br />

During the process of casting, a viscous oxide film forms on the surface of the molten<br />

TABLE 10-7 Composition of 16% High-Chrome Abrasion-Resistant White Iron<br />

France Germany United Kingdom United States<br />

Standard NF 32-401<br />

(1980)<br />

DIN 1695<br />

(1981)<br />

BS 4844 Pt2:1972 ASTM A532 1982<br />

Grade FB Cr9 Ni5 GX 300 CrMo 3A 3B Class II Class II<br />

15 3 Ty<strong>pe</strong> B Ty<strong>pe</strong> C<br />

C (%) 2.0–3.6 2.3–3.6 2.4–3.0 3.0–3.6 2.4–2.8 2.8–3.6<br />

Si (%) 0.2–0.8 0.2–0.8 1.0 max 1.0 max 1.0 max 1.0 max<br />

Mn (%) 0.5–1.0 0.5–1.0 0.5–1.0 0.5–1.0 0.5–1.0 0.5–1.0<br />

Ni (%) 0.0–2.5 0.7 0.0–1.0 0.0–1.0 0.5 max 0.5 max<br />

Cr (%) 14.0–17.0 14.0–17.0 14.0–17.0 14.0–17.0 14.0–18.0 14.0–18.0<br />

Cu (%) — 0–1.2 0–1.2 1.2 max 1.2 max<br />

Mo (%) 0.5–3.0 1.0–3.0 0.0–2.5 1.0–3.0 1.0–3.0 2.3–3.5<br />

P max (%) — 0.3 0.3 0.1 0.1<br />

S max (%) — — 0.1 0.06 0.06<br />

Data from Brown (1994).


MATERIALS SCIENCE FOR SLURRY SYSTEMS<br />

TABLE 10-8 Composition of 27% High-Chrome Abrasion-Resistant White Iron<br />

10.9<br />

France Germany United Kingdom United States<br />

Standard NF 32-401<br />

(1980)<br />

DIN 1695 (1981) BS 4844 Pt2:1972<br />

ASTM A532<br />

(1982)<br />

Grade FB Cr26MoNI GX 260 Cr27 GX300CrMo 3D 3E Class III<br />

27 1 Ty<strong>pe</strong> A<br />

C (%) 1.5–3.5 2.3–2.9 3.0–3.5 2.4–2.8 2.8–3.2 2.3–3.0<br />

Si (%) 0.2–1.2 0.5–1.5 0.2–1.0 1.0 max 1.0 max 0.5–1.5<br />

Mn (%) 0.5–1.5 0.5–1.5 0.5–1.0 0.5–1.5 0.5–1.5 1.0 max<br />

Ni (%) 0.0–2.5 1.2 1.2 0.0–1.0 0.0–1.0 1.5 max<br />

Cr (%) 22.0–28.0 24.0–28.0 23.0–28.0 22.0–28.0 22.0–28.0 23.0–28.0<br />

Cu (%) 0.0–1.5 2.0 max 2.0ax 1.2 max<br />

Mo (%) 0.5–3.0 1.0 1.0–2.0 0–1.5 0–1.5 1.5 max<br />

P (%) — 0.1 0.1 0.1<br />

S (%) — — 0.1 0.06 0.06<br />

Data from Brown (1994).<br />

metal in the furnace or ladle. It must be skimmed before pouring and the running system<br />

must be designed to trap oxide dross. Metal filters are used wherever possible. Gating<br />

should be designed to provide rapid filling of the mold with minimum turbulence.<br />

The alloy shown in Table 10-7 is sometimes called 15/3 Chrome/Moly Iron, or 16%<br />

chrome. It is a martensitic white iron of moderate erosion resistance. It is used for the<br />

casting of pumps for a number of applications such as carbon in pulp circuits, coal transfer<br />

pumps, sewage treatment, and some newspa<strong>pe</strong>r recycling pumps.<br />

The alloy known in the industry as 20% chrome, also known as ASTM A532-87 Class<br />

II Ty<strong>pe</strong> D, is available with a tensile strength of 595 to 875 MPa and with suitable multistep<br />

heat treatment can have a minimum hardness of 650 HBN or 59 HRC. It is used for<br />

the manufacture of mill liners, <strong>slurry</strong> pump parts, pulverizer parts, and other wear-resistant<br />

castings. Its composition includes carbon in the range of 2 to 3.3%, chromium in the<br />

range of 18 to 23%, and molybdenum at a maximum of 3%. The structure consists of eutectic<br />

and secondary carbides in a martensitic matrix.<br />

The alloy known in the industry as 27% chrome, whose composition is shown in<br />

Table 10-8, is very popular with manufacturers of <strong>slurry</strong> pumps, as it offers abrasion resistance,<br />

erosion resistance, mild corrosion resistance, and mild heat resistance. It consists<br />

of a microstructure of chromium carbides in a martensitic–austenitic matrix. Its tensile<br />

strength is in the range of 525 to 700 MPa (75,000–100,000 psi) and can be heat treated to<br />

a minimum hardness of 650 HBN/58 HRC. Table 10-9 summarizes the classes of white<br />

iron.<br />

Corrosion can have very detrimental effects on the wear resistance of certain white<br />

irons. If the matrix wears out and exposes the carbides to a corrosive environment, they<br />

may quickly deteriorate, causing a further loss of wear resistance. Manufacturers of <strong>slurry</strong><br />

pumps have gone beyond the s<strong>pe</strong>cifications of ASTM A532-87 to increase the chromium<br />

contents to 30% with 50% chromium carbides in a martensitic matrix. These materials<br />

have found applications in phosphate matrix and phosphoric acid pumping. These new alloys<br />

are intermediary between white irons and su<strong>pe</strong>ralloys. Walker (1990) reported that<br />

the use of these alloys leads to a substantial increase of the wear life over conventional<br />

white iron castings (Ni-hard, 27% high chrome).<br />

The corrosion pro<strong>pe</strong>rties of stainless steel and duplex stainless are enhanced by adding<br />

high-chromium carbides in the matrix to produce a range of materials suitable for flue gas<br />

desulfurization. The matrix of the alloy is a balance between ferritic stainless steel rich in


10.10 CHAPTER TEN<br />

TABLE 10-9 Comparison of White Iron Classes<br />

Hypoeutectic Eutectic Hy<strong>pe</strong>reutectic<br />

Example Ni-hard 27% chrome Very high chrome (>30%)<br />

Matrix and carbides Soft primary ferrous White iron matrix with White iron austenitic/<br />

dendrites with eutectic very fine dis<strong>pe</strong>rsion of martensite matrix with<br />

carbides eutectic carbides coarse, discrete primary<br />

carbides<br />

Note on casting Refer to ASTM 532 Solidification during<br />

for maximum sizes of casting must be controlled<br />

castings to avoid cracking<br />

soluble chromium, nickel, and other alloys to resist corrosion and a substantial volume of<br />

chromium carbides to resist erosion wear.<br />

When designing components of pumps subject to torsion and bending loads, such as<br />

shaft sleeves and pump casings, it is important to appreciate that the harder the material,<br />

the narrower the limit on strain or elongation. A range of elongation of 0.5% to 1.5% is a<br />

minimum for bowls, mantle liners, pump-wetted parts, ball mill grates, and components<br />

subject to impact loads.<br />

The author was once asked to express an opinion on the failure of a large number of<br />

stainless steel shaft sleeves at start-up. The <strong>slurry</strong> pumps used very hard shaft sleeves.<br />

Upon investigation it was noticed that the crack was longitudinal. A modeling of the shaft<br />

sleeve by finite element analysis revealed that the strain due to torsion exceeded the allowed<br />

maximum value. These large pumps were using motors with 250% starting torque.<br />

That initial torque screwed the pump im<strong>pe</strong>ller tightly against the shaft sleeve and applied<br />

FIGURE 10-3 Mictograph showing grain structure of Moballoy, a proprietary high-chrome,<br />

abrasion-resisting alloy. (Courtesy of Mobile Pulley and Machine Works, Inc.)


a strong torsion moment. The shaft sleeve, due to its excessive hardness, simply cracked.<br />

It had to be replaced by shaft sleeves with a lower hardness. The manufacturer eventually<br />

develo<strong>pe</strong>d a s<strong>pe</strong>cial shaft sleeve with a hard ceramic face, but with a more ductile base to<br />

take the torsion loads.<br />

The resistance of the matrix and carbides to wear is a complex phenomenon due to<br />

hardness of the individual components. Ferrite matrix has a mere hardness of 150–250<br />

HV. Austenite, as found in many stainless steels, has a hardness of 300–500 HV. Martensite,<br />

which is sometimes obtained by very fast solidification or chilling of the casting, has<br />

a hardness of 500–1000 HV. The iron carbides Fe 3C have a hardness in the range of<br />

850–1000 HV, Chromium carbide (FeCr) 7C 3 has a hardness of 1400–1600 HVas primary<br />

carbides and a hardness of 1200–1400 as eutectics (Huggett and Walker, 1992).<br />

Dredge pumps and ball mill liners must be able to absorb the energy of impact without<br />

fracture. The energy due to impact involves both loads and deflections. The capacity to absorb<br />

such energy is called the resilience. Its measure is the modulus of resilience. For steels,<br />

it is the area under the stress-to-strain triangle in the elastic range shown in Figure 10-1.<br />

The fracture toughness is a relative measure of the ability to absorb shock loads. This<br />

is particularly important with dredge pumps. Certain 16% high-chrome alloys (ASTM<br />

A532 Class II Ty<strong>pe</strong> B) have a fracture toughness in the range of 14–28 KSI-in. 20% highchrome<br />

alloys (ASTM A532 Class II Ty<strong>pe</strong> E) have a fracture toughness in the range of<br />

21–28 KSI-in. 27% high chrome alloys (ASTM A532 Class III Ty<strong>pe</strong> A) have a fracture<br />

toughness in the range of 22–27 KSI-in. Because different manufacturers and foundries<br />

apply different levels of heat treatments, these values should be taken as indicative. The<br />

manufacturer should always be consulted.<br />

10-4 NATURAL RUBBERS<br />

MATERIALS SCIENCE FOR SLURRY SYSTEMS<br />

10.11<br />

Elastomers, particularly different grades of rubber, are widely used in the lining of <strong>slurry</strong><br />

handling equipment as well as piping. Natural rubber is preferred for solids with a diameter<br />

smaller than 6 mm or 1 – 4 in for pump im<strong>pe</strong>llers and smaller than 15 mm or 5 – 8 in for pump<br />

liners. Natural rubber resists well a number of acids with a pH less than 6.0, but is not<br />

suitable for pumping oils or solvents. Liquid tem<strong>pe</strong>rature must be lower than 65°C, or<br />

150°F.<br />

The abrasion resistance of natural rubber exceeds the resistance of all other elastomers.<br />

A number of forms of natural rubber are available such as:<br />

TABLE 10-10 Pro<strong>pe</strong>rties of Natural Aashto<br />

Durometer (Shore A) hardness 50 (±5) to 70 (±5)<br />

Tensile strength (MPa) 22<br />

Tensile strength (psi) 3190<br />

Elongation (%) 450<br />

S<strong>pe</strong>cific gravity 1.07<br />

Tear resistance Excellent<br />

Abrasion resistance Good<br />

Impact resistance Excellent<br />

Flame resistance Poor<br />

Heat aging at 100°C (212°F) Average<br />

Weather resistance Good


10.12 CHAPTER TEN<br />

� Pure tan gum natural rubber<br />

� Natural aashto<br />

� Carbon-filled natural rubber<br />

� White food-grade natural rubber<br />

� Carbon-filled and silica-filled natural rubber<br />

� Hard natural rubber/butadiene styrene compound filled with graphite<br />

10-4-1 Natural Aashto<br />

Natural aashto is a natural rubber of excellent resistance to impact and to tear, with good<br />

resistance to abrasion. It is weather-resistant and is widely used as a certified bridge construction<br />

material. It is available in 50, 60, and 70 durometer hardness. Its tem<strong>pe</strong>rature<br />

range for use is from –50°C to 100°C (–60°F to 212°F). The mechanical pro<strong>pe</strong>rties of natural<br />

aashto are listed in Table 10-10.<br />

10-4-2 Pure Tan Gum<br />

Pure tan gum is a natural rubber of excellent resistance to abrasion and high tensile<br />

strength. It is excellent as protection against impinging abrasion in chutes, troughs, and<br />

hop<strong>pe</strong>rs. It is available in 35 to 45 durometer shore A hardness. Its tem<strong>pe</strong>rature range for<br />

use is from –50°C to 65°C (–60°F to 150°F). It has poor resistance to ozone. The mechanical<br />

pro<strong>pe</strong>rties of pure tan gum are listed in Table 10-11.<br />

10-4-3 White Food-Grade Natural Rubber<br />

A s<strong>pe</strong>cial high-quality odorless natural rubber is available with a Shore A hardness of 40.<br />

It is limited to a tem<strong>pe</strong>rature of 65°C (150°F).<br />

TABLE 10-11 Pro<strong>pe</strong>rties of Pure Tan Gum<br />

Durometer (Shore A) hardness 40 (±5)<br />

Tensile strength (MPa) 17<br />

Tensile strength (psi) 2500<br />

Elongation (%) 600<br />

Tear resistance Excellent<br />

Abrasion resistance Excellent<br />

Impact resistance Excellent<br />

Flame resistance Poor<br />

Ozone resistance Poor<br />

Heat aging at 100°C (212°F) Good<br />

Resistance to oil and <strong>pe</strong>troleum products Poor, oil and <strong>pe</strong>troleum products cause swelling<br />

and degradation<br />

Typical applications Sand and gravel<br />

Tip s<strong>pe</strong>ed limit for pump im<strong>pe</strong>llers 25 m/s (4920 ft/min)


10-4-4 Carbon-Black-Filled Natural Rubber<br />

This is an industrial grade of natural rubber with better tear resistance but slightly lower<br />

abrasion resistance than natural rubber. It typically has durometer hardness 45-Shore<br />

hardness A. Its tem<strong>pe</strong>rature limit is 82°C (180°F). This form of filled natural rubber is<br />

widely used in the industrial, mining, and utilities industries.<br />

10-4-5 Carbon-Black- and Silicon-Filled Natural Rubber<br />

These fillers are added to natural rubber to slightly improve the resistance to traces of oil<br />

and to stabilize the matrix against thermal degradation by friction. It typically has durometer<br />

hardness 55 Shore hardness A. Its tem<strong>pe</strong>rature limit is 82°C (180°F). The tip s<strong>pe</strong>ed of<br />

the im<strong>pe</strong>ller can be a maximum of 28 m/s (5500 ft/sec).<br />

10-4-6 Hard Natural Rubber/ Butadiene Styrene Compound Filled<br />

with Graphite<br />

This compound is resistant to chlorine and a wide range of acids. It is particularly hard after<br />

curing, which allows machining. Its tem<strong>pe</strong>rature limit is 93°C (200°F). Nominal<br />

durometer hardness is 60 Shore D.<br />

Rubber-lined products that are subject to long storage must be stabilized against storage-based<br />

degradation by suitable antioxidants and antidegradants.<br />

10-5 SYNTHETIC RUBBERS<br />

In addition to natural rubber, a number of synthetic elastomers are available for handling<br />

slurries.<br />

10-5-1 Polychlorene (Neoprene)<br />

MATERIALS SCIENCE FOR SLURRY SYSTEMS<br />

10.13<br />

Neoprene, a dupont trademark for polychlorene, is su<strong>pe</strong>rior to natural rubber in its resistance<br />

to oils and chemicals, and for higher-tem<strong>pe</strong>rature applications. Although Neoprene<br />

TABLE 10-12 Pro<strong>pe</strong>rties of Polychlorene (Neoprene)<br />

Durometer (Shore A) hardness 40 ± 5 50 ± 5 60 ± 5 70 ± 5 80 ± 5<br />

Tensile strength (MPa) 8 8 8 8 8<br />

Tensile strength (psi) 1160 1160 1160 1160 1160<br />

Elongation (%) 560 500 350 250 200<br />

Tear resistance Good<br />

Abrasion resistance Good<br />

Impact resistance Good<br />

Flame resistance Good<br />

Ozone resistance Good<br />

Typical tem<strong>pe</strong>rature range –34°C to 100°C<br />

–30°F to 212°F


10.14 CHAPTER TEN<br />

TABLE 10-13 Pro<strong>pe</strong>rties of EPDM<br />

Durometer (Shore A) Hardness 60 ± 5<br />

Tensile strength (MPa) 7<br />

Tensile strength (psi) 1000<br />

Elongation (%) 400<br />

Tear resistance Average<br />

Abrasion resistance Good<br />

Impact resistance Good<br />

Flame resistance Poor<br />

Ozone resistance Excellent<br />

Weather resistance Excellent<br />

Typical tem<strong>pe</strong>rature range –50°C to 177°C<br />

–60°F to 350°F<br />

has a lower resistance to abrasion than narural rubber, it is the next most common material<br />

for pump parts. The tip s<strong>pe</strong>ed for im<strong>pe</strong>llers is 33 m/s (or 6496 ft/sec). The mechanical<br />

pro<strong>pe</strong>rties of Neoprene in different hardness grades are presented in Table 10-12.<br />

A form of hard Neoprene is available for up<strong>pe</strong>r and lower screens of vertical sump<br />

pumps and for cap nuts and flingers that protect the shaft seals. Its nominal hardness is 80<br />

Shore A, but it can be postcured to a hardness of 85 Shore A.<br />

Because of its su<strong>pe</strong>rior resistance to oils and hydrocarbons, Neoprene is preferred to<br />

natural rubber for pumping oil (tar) sand slurries, oil shales, pulp and pa<strong>pe</strong>r, and phosphate<br />

and in moly circuits when oils are used as flotation reagents. A s<strong>pe</strong>cial curing<br />

process is applied when molding equipment parts out of Neoprene to give it excellent resistance<br />

to water absorption and swelling, and suitable thermal resistance for use up to a<br />

tem<strong>pe</strong>rature of 120°C (248°F) with a nominal hardness of 60 Shore A.<br />

Neoprene is affected or attacked by strong oxidizing acids, acetic acid, ketones, esters,<br />

and chlorinated and nitro hydrocarbons.<br />

TABLE 10-14 Pro<strong>pe</strong>rties of Jade Green Aramound<br />

Durometer (Shore A) Hardness 60 ± 5<br />

Tensile strength (MPa) 22<br />

Tensile strength (psi) 3190<br />

Elongation (%) 700<br />

Tear resistance Excellent<br />

Abrasion resistance Excellent<br />

Impact resistance Excellent<br />

Flame resistance Poor<br />

Ozone resistance Poor<br />

Weather resistance Average<br />

Typical tem<strong>pe</strong>rature range –50°C to 66°C<br />

–60°F to 150°F


10-5-2 Ethylene Propylene Terpolymer (EPDM)<br />

EPDM has a generally acceptable resistance to most moderate chemicals, alcohol, ozone,<br />

and organic acids. It is, however, attacked by strong acids, solvents, most hydrocarbons,<br />

chloroform, and aromatic solvents. The minimum tem<strong>pe</strong>rature for use is –51°C (–60°F)<br />

and its maximum tem<strong>pe</strong>rature is 177°C or 350°F. EPDM offers better abrasion resistance<br />

than butyl rubber. It is also used as a high-tem<strong>pe</strong>rature gasket. Mechanical pro<strong>pe</strong>rties are<br />

presented in Table 10-13.<br />

10-5-3 Jade Green Armabond<br />

This product is highly resilient and abrasion resistant. It is a used as a soft facing and protective<br />

lining for material handling equipment such as chutes, launders, ball mills, and cement<br />

mixers. Goodyear manufactures this material. Its mechanical pro<strong>pe</strong>rties are presented<br />

in Table 10-14.<br />

10-5-4 Armadillo<br />

Armadillo is a synthetic rubber produced by Goodyear. It is highly resistant to abrasion. It<br />

is used to line chutes, launders, and troughs and provides excellent noise abatument. The<br />

material may be supplied with a heavy nylon backing to be self-supporting and bridge<br />

gaps in chutes and launders. It has a nominal hardness 60 shore A.<br />

10-5-5 Nitrile<br />

MATERIALS SCIENCE FOR SLURRY SYSTEMS<br />

10.15<br />

Nitrile offers moderate resistance to abrasion with excellent ability to handle oils and hydrocarbons.<br />

A s<strong>pe</strong>cial food grade is available for the food and beverage industry. The<br />

pro<strong>pe</strong>rties of nitrile are presented in Table 10-15. Nitrile may be reinforced with heavy<br />

nylon fabric for high-pressure hoses. Nitrile is used for seals, flingers, and screens. It is<br />

attacked by ozone, ketones, esters, aldehydes, and chlorinated and nitro-hydrocarbons.<br />

TABLE 10-15 Pro<strong>pe</strong>rties of Nitrile—Food Grade<br />

Durometer (Shore A) Hardness 60 ± 5<br />

Tensile strength (MPa) 5<br />

Tensile strength (psi) 700<br />

Elongation (%) 450<br />

Tear resistance Good<br />

Abrasion resistance Good<br />

Impact resistance Fair<br />

Flame resistance Poor<br />

Gas im<strong>pe</strong>rmeability Good<br />

Typical tem<strong>pe</strong>rature range –35°C to 121°C<br />

–30°F to 250°F


10.16 CHAPTER TEN<br />

TABLE 10-16 Pro<strong>pe</strong>rties of Hypalon—Food Grade<br />

Durometer (Shore A) Hardness 60 ± 5<br />

Tensile strength (MPa) 10<br />

Tensile strength (psi) 1400<br />

Elongation (%) 500<br />

Tear resistance Average<br />

Abrasion resistance Good<br />

Impact resistance Good<br />

Flame resistance Good<br />

Gas im<strong>pe</strong>rmeability Good<br />

Heat aging Good<br />

Weather resistance Excellent<br />

Ozone resistance Excellent<br />

TABLE 10-17 Pro<strong>pe</strong>rties of Thermoset Polyurethane<br />

Casting resins<br />

50–60% Mineral filled<br />

potting and casting<br />

Pro<strong>pe</strong>rties Liquid Unsaturated compounds<br />

Processing tem<strong>pe</strong>rature 85–121°C 121°C–250°F<br />

185–250°F (casting)<br />

Molding pressure 0.7–35 MPa<br />

100–5000 psi<br />

Mold (Linear) shrinkage (in/in) 0.020 0.001–0.002<br />

Tensile strength at break 1.2–69 MPa 69–76 MPa 6.9–48.3 MPa<br />

0.175–10 ksi 10–11 ksi 1–7 ksi<br />

Elongation at break (%)<br />

Compressive strength<br />

100–1000 3–6 5–55<br />

(yield/rupture) 138 MPa<br />

20,000 psi<br />

Flexural strength 4.8–31 MPa 131 MPa<br />

(yield/rupture) 700–4,500 psi 19,000 psi<br />

Tensile modulus 69–689 kPa<br />

10–100 ksi<br />

Compressive modulus 69–689 kPa<br />

10–100 ksi<br />

Flexural modulus at 23°C, 73°F 69–689 kPa 4200 kPa<br />

10–100 ksi 610<br />

Hardness Shore A10 Barcol 30-35 Shore A90<br />

D 90 D 52-85<br />

S<strong>pe</strong>cific gravity 1.03–1.5 1.05 1.37–2.1<br />

Water absorption over 24 hr in<br />

a 3.2 mm (1/8 in) s<strong>pe</strong>cimen<br />

0.2–1.5% 0.1–0.2% 1.37–2.1%<br />

Data from Har<strong>pe</strong>r, 1992.


10-5-6 Carboxylic Nitrile<br />

Carboxylic nitrile combines excellent oil resistance with abrasion resistance. It is available<br />

in nominal hardness of 80 Shore A. It exhibits a high tensile modulus, low elongation,<br />

improved hot tear and tensile resistance, and better resistance to hot oil and air aging<br />

than most nitriles.<br />

10-5-7 Hypalon<br />

MATERIALS SCIENCE FOR SLURRY SYSTEMS<br />

10.17<br />

Hypalon is a tradename for CSM chlorosulfonated polyethylene. It offers good resistance<br />

to moderate chemicals, ozone, alkaline solutions, hydrogen, Freon, alcohols, aliphatic hydrocarbons,<br />

as well as ultraviolet degradation from sunrays. Strong oxidizing acids, ketones,<br />

esters, acetic acid, and chlorinated and nitro-hydrocarbons attack Hypalon. Its tem<strong>pe</strong>rature<br />

range for applications is from –40°C to 150°C (–40°F to 300°F). Its physical<br />

pro<strong>pe</strong>rties are presented in Table 10-16.<br />

TABLE 10-18 Pro<strong>pe</strong>rties of Thermoplastic Polyurethane<br />

Unreinforced<br />

10–20%<br />

glass-fiber-reinforced Long glass-reinforced<br />

Pro<strong>pe</strong>rties molding compound molding compounds molding compound<br />

Melting tem<strong>pe</strong>rature 24–58°C<br />

75–137°F<br />

75°F<br />

Processing tem<strong>pe</strong>rature 221–266°C 182–210°C 182–232°F<br />

430–510°F 360–410°F 360–450°F<br />

Molding pressure 55–76 MPa<br />

8–11 ksi<br />

Tensile strength at break 50–62 MPa 69–76 MPa 186–227 MPa<br />

7.2–9 ksi 4.8– 7.5 ksi 27–33 ksi<br />

Elongation at break (%) 60–180 3–70 2<br />

Tensile yield strength, psi 53.7–76 MPa 186–227 MPa<br />

7.8–11.0 ksi 27–33 ksi<br />

Compressive strength 35 MPa<br />

(yield/rupture) 5000 psi<br />

Flexural strength 4.8–31 MPa 11.7–42 MPa 310–393 MPa<br />

(yield/rupture) 10.2–15 ksi 1.7–6.2 ksi 45–57 ksi<br />

Tensile modulus 1.31–2.07 GPa 4.1–9.6 MPa 11.7–17.2 GPa<br />

190–300 ksi 0.6–1.40 ksi 1700–2500 ksi<br />

Flexural modulus at 1.62–2.14 GPa 0.28–0.62 GPa 10.3–15.16 GPa<br />

23°C, 73°F 235–310 ksi 40–90 ksi 1500–2200 ksi<br />

Hardness R > 100, M48 R 45-55<br />

S<strong>pe</strong>cific gravity 1.2 1.22–1.38<br />

Water absorption over 24 0.17–0.19% 0.4–0.55%<br />

hr in a 3.2 mm (1/8 in)<br />

s<strong>pe</strong>cimen<br />

Data from Har<strong>pe</strong>r, 1992.


10.18 CHAPTER TEN<br />

10-5-8 Fluoro-elastomer (Viton)<br />

This elastomer offers exceptional resistance to oils and many chemicals at elevated tem<strong>pe</strong>rature,<br />

but limited resistance to erosion wear.<br />

10-5-9 Polyurethane<br />

Polyurethane is a castable synthetic rubber. It is elected for applications where tramp is a<br />

problem. It offers excellent resistance to high tear and high tensile strength. Its resistance<br />

to erosion is, however, lower than natural rubber.<br />

Polyurethane pump liners and pump parts are selected for fine slurries with an average<br />

particle diameter of (minus 230 (m, minus 65 mesh). Im<strong>pe</strong>llers may o<strong>pe</strong>rate up to a tip<br />

s<strong>pe</strong>ed of 31 m/s (6100 ft/sec). Polyurethane lined pumps are used for flue gas desulfurization<br />

applications as well as for pumping fairly fine tailings from cop<strong>pe</strong>r and gold plants<br />

and for moly and flotation circuits where oils are used as reagents. The up<strong>pe</strong>r limit for use<br />

of polyurethane is 82°C. Nominal hardness is 80 Shore A.<br />

Polyurethane is subject to hydrolic degradation at its higher tem<strong>pe</strong>rature limits.<br />

Polyurethane is machinable. It does not seal as well as rubber, and additional O-rings and<br />

gaskets are needed when switching from rubber to polyurethane lined parts. Polyurethane<br />

is available in thermoset and thermoplastic forms. A new generation of glass-reinforced<br />

polyurethane thermoplastic materials are available from different manufacturers. The<br />

pro<strong>pe</strong>rties of polyurethane are presented in Tables 10-17 and 10-18.<br />

10-6 WEAR DUE TO SLURRIES<br />

A predominant factor in wear life and the selection of the pro<strong>pe</strong>r material for the manufacture<br />

of <strong>slurry</strong> handling equipment and piping is the abrasiveness of slurries. This is often<br />

related to the hardness of the particles, their degree of roundness, their concentration,<br />

the angle of impingement, and corrosion. Wilson (1985) plotted the hardness of different<br />

materials on a scale (Figure 10-2) and develo<strong>pe</strong>d a chart for the selection of materials for<br />

the construction of centrifugal pumps (Table 10-19). Metals with high elastic limits are<br />

used to absorb abrasion associated with impact. Hard metals are particularly brittle and<br />

are selected where there is low impact but relatively hard particles flowing in parallel to<br />

the surface of the equipment.<br />

The rate of wear or mass loss is a function of the s<strong>pe</strong>ed of the flow raised to the power<br />

of 2.5 to 4.0 (Wilson, 1985). Wear is also accentuated when the solids are trap<strong>pe</strong>d between<br />

a rotating and a fixed surface. This occurs, for example, between the back wear<br />

plate and the im<strong>pe</strong>ller (Figure 10-4).<br />

The American Society for Testing of Materials (Miller, 1987) has adopted the concept<br />

of the Miller number as a measure of the abrasivity of slurries. Standard G-75 describes<br />

the relative abrasivity based on the mass loss of a standard 27% chrome white<br />

iron block when rubbed in <strong>slurry</strong> for a particular <strong>pe</strong>riod of time. To conduct the test and<br />

to measure the <strong>slurry</strong> abrasion response (SAR) number, also called Miller number, a<br />

s<strong>pe</strong>cial machine (Figure 10-5) is used. It consists of a wear block driven at 48 strokes<br />

<strong>pe</strong>r minute with a stroke length of 200 mm (8 in). A dead weight of 2.27 kg [5 lb] is<br />

applied to the wear block. At the bottom of the tray, a piece of Neoprene is installed to<br />

act as a lap. The tray takes the form of a flat bottom between V-sha<strong>pe</strong>d walls that confine<br />

the <strong>slurry</strong> around the block. The tray is filled with sand. The mounted blocks are<br />

lowered in the trays. Four, 4-hour tests are conduced and the weight of the block is


MATERIALS SCIENCE FOR SLURRY SYSTEMS<br />

TABLE 10-19 Classification of Pumps According to Solid Particle Size<br />

a Theoretical value. From Pump Handbook, reprinted by <strong>pe</strong>rmission of McGraw-Hill.<br />

10.19


10.20 CHAPTER TEN<br />

FIGURE 10-4 Wear for the wear plate of a concentrate charge pump.<br />

measured after each run to obtain the mass loss as a form of wear. The data is plotted<br />

to obtain the following correlation between mass loss and time:<br />

mass loss = A·tB where<br />

t = time<br />

A = coefficient<br />

B = power number<br />

The Miller is determined at the end of the first 2 hours of the test as<br />

Miller number = 18.18 [A·B· 2B–1 ] (10-6)<br />

The Miller number is directly related to the abrasivity of the <strong>slurry</strong>. In other words, <strong>slurry</strong><br />

FIGURE 10-5 Machine for measuring the Miller number. (From Miller, 1987. Reproduced<br />

by <strong>pe</strong>rmission of ASTM.)


with a Miller number of 150 will be twice as destructive as <strong>slurry</strong> with a Miller number of<br />

75. Examples of Miller numbers are presented in Table 10-20. Typically, slurries with a<br />

Miller number smaller than 50 are considered to cause mild wear. The Miller number is<br />

also a function of the weight concentration (Figure 10-6).<br />

The ASTM has develo<strong>pe</strong>d a new standard (G75-01) to determine the <strong>slurry</strong> abrasivity<br />

(Miller) number and the <strong>slurry</strong> response of materials (SAR number).<br />

10-7 CONCLUSION<br />

MATERIALS SCIENCE FOR SLURRY SYSTEMS<br />

TABLE 10-20 Miller Number of Certain Slurries<br />

Material Miller number<br />

Alundum, 400 mesh 241<br />

Alundum, 200 mesh 1058<br />

Ash 127<br />

Fly ash 83, 14<br />

Bauxite 9, 33, 50, 134<br />

Calcium carbonate 14<br />

Carbon 16<br />

Carborundum, 220 mesh 1284<br />

Clay 36<br />

Coal 6, 10, 21, 28, 57<br />

Cop<strong>pe</strong>r concentrate 19, 37, 68, 128<br />

Dust, blast furnace 57<br />

Gypsum 41<br />

Iron ore (or concentrate) 28, 64, 122, 234<br />

Kaolin 7, 30<br />

Limestone 22, 39, 46<br />

Limonite 113<br />

Magnetite 64, 71, 134<br />

Mud, drilling 10<br />

Phosphate 68, 74, 84, 134<br />

Pyrite 194<br />

Sand, silica 51, 68, 116, 246<br />

Sewage (digested) 15<br />

Sewage (raw) 25<br />

Sulfur 1<br />

Tailings (all ty<strong>pe</strong>s) 24, 91, 217, 644<br />

Data from Miller, 1987.<br />

10.21<br />

A variety of materials are available for the lining and manufacture of <strong>slurry</strong> handling<br />

equipment and piping. Various selections are available based on the degree of abrasiveness,<br />

corrosion, tem<strong>pe</strong>rature, and presence of oils or solvents. In terms of measurement of<br />

wear, the American Society for Testing of Materials has develo<strong>pe</strong>d a <strong>slurry</strong> abrasion response<br />

(SAR) number, also called Miller number. It is extremely useful for determining<br />

the need to line pi<strong>pe</strong>lines and pumps.


10.22 CHAPTER TEN<br />

FIGURE 10-6 The Miller number as a function of the weight concentration of solids. This<br />

example uses 70 mesh sand. (From Miller, 1987. Reproducted by <strong>pe</strong>rmission of ASTM.)<br />

REFERENCES<br />

ASTM. 1999. ASTM A532/A532M. Standard S<strong>pe</strong>cification for Abrasion-Resistant Cast Irons.<br />

ASTM. 2001. Test Method G75-01. Standard Test Method for Determination of Slurry Abrasivity<br />

(Miller Number) and Slurry Response of Materials (SAR Number).<br />

BSI. 1986—withdrawn. S<strong>pe</strong>cification for Abrasion-Resisting White Cast iron.<br />

Brown, J. R. 1994. FOSECO Foundryman’s Handbook. London: Butterworth Heinemann.<br />

Farag, M. M. 1997. Minerals Selection for Engineering Design. Up<strong>pe</strong>r Saddle River, NJ: Prentice<br />

Hall.<br />

Har<strong>pe</strong>r, A. C. 1992. Handbook of Plastics, Elastomers, and Composites. New York: McGraw-Hill.<br />

Juvinall, R. C. 1983. Fundamentals of Machine Component Design. New York: Wiley.<br />

Miller, J. E. 1987. Slurry Abrasivity Determination. The American Society for Testing of Materials<br />

Standardization News, July.<br />

Wilson, G. 1985. Construction of Solids-handling Centrifugal Pumps. Chapter 9-17-2 in The Pump<br />

Handbook, I. J. Karassik, W. C. Krutzch, W. H. Fraser, and J. P. Messina Eds. New York:<br />

McGraw-Hill.<br />

Zakharov, B. 1962. Heat Treatment of Metals. Moscow: Peace Publishers.<br />

Further readings. The following technical bulletins may be obtained directly from Weir Slurry<br />

Group for further readings on white chrome irons:<br />

Dolman, K. F. 1992. Ultrachrome—Corrosion Resistant Alloys. Warman Technical Bulletin No. 19,<br />

Sydney, Australia.<br />

Huggett, P. G. and Walker, C. I. 1992. White Iron Microstructure and Wear. Warman Technical Bulletin<br />

No 18, Sydney, Australia.<br />

Walker, C. I. 1990. Hy<strong>pe</strong>rchrome White Irons. Warman Technical Bulletin No 4, Sydney, Australia.


CHAPTER 2<br />

FUNDAMENTALS OF<br />

WATER FLOWS IN PIPES<br />

2-0 INTRODUCTION<br />

The mechanics of pi<strong>pe</strong> flow is a topic well dealt with in the scientific literature. In this<br />

chapter, some of the concepts are reviewed in a simplified manner as they relate to <strong>slurry</strong><br />

flows.<br />

The friction factor for single phase flows is presented in terms of boundary layer theory.<br />

Losses in pi<strong>pe</strong>s are summarized in terms of fittings and conduits commonly used on<br />

large engineering projects. Numerous books have been written for single-phase flows.<br />

This chapter limits itself to a brief introduction.<br />

The equations in this chapter are based on SI units for consistency. They can be readily<br />

used in USCS (United States Customary System) units if the reader assumes the use of<br />

slugs and not pounds for mass units. There is considerable confusion about the use of<br />

slugs, which are units of mass, and pounds, which are units of force. The conversion factor<br />

between these is often called g c and is equal to 32.2 ft/sec. Many other references input<br />

g c in their equations to achieve such a conversion, but it was not deemed necessary in<br />

this book. The worked examples show how slugs should be used as a unit of mass.<br />

Certain models of Newtonian <strong>slurry</strong> flows are attempts to apply correction factors to<br />

the friction losses of the carrier liquid on the basis of the volumetric concentration of<br />

solids. The reader is encouraged to examine this chapter before proceeding with more<br />

complex themes.<br />

2-1 SHEAR STRESS OF LIQUID FLOWS<br />

Modern fluid mechanics is based on the concept of a controlled volume. Mass momentum<br />

and energy must be conserved when a particle enters and leaves the volume.<br />

Considering flow through a section of pi<strong>pe</strong> of a constant diameter between two locations<br />

1 and 2 as in Figure 2-1, the hydraulics force associated with the drop of pressure<br />

is<br />

�<br />

F12 = � 2 D i (P1 – P2) (2-1)<br />

4<br />

2.1


2.2 CHAPTER TWO<br />

This force is balanced by the friction force F r<br />

F r = � w�D iL (2-2)<br />

where L is the distance between points 1 and 2 and � w is the wall shear stress<br />

or<br />

P<br />

1<br />

�<br />

� 4<br />

D i 2 (P1 – P 2) – � w�DL = 0<br />

Di(P1 – P2) Ri(P1 – P2) �w = �� = ��<br />

(2-3)<br />

4L 2L<br />

The shear stress at any radius and from the center of the pi<strong>pe</strong> is<br />

Ri�P r<br />

� = � = �w �<br />

2L R<br />

where<br />

L = length<br />

Ri = the pi<strong>pe</strong> inner radius (at the inside wall of the pi<strong>pe</strong>)<br />

r = local radius<br />

The shear stress � is calculated. At the center of the pi<strong>pe</strong> there is no shear stress.<br />

L<br />

Example 2-1<br />

Homogeneous <strong>slurry</strong> is tested in a pi<strong>pe</strong> with an inner diameter of 53 mm (2.086 in). The<br />

pressure drop due to friction is measured between two points (A and B), which are separated<br />

by a distance of 1.8 m (5.9 ft). The pressure drop is recorded as 3000 Pa (0.435 psi).<br />

To appreciate the shear stress distribution from the wall to the center of the pi<strong>pe</strong>, determine<br />

the shear stress distribution at the wall and at three points: at a radius of 20 mm<br />

(0.787�), at a radius of 12 mm (0.472�), and at the center of the pi<strong>pe</strong>.<br />

Solution<br />

This problem will be solved in SI units [Système International (metric)] and in USCS<br />

units.<br />

w<br />

P<br />

2<br />

FIGURE 2-1 Shear stress and pressure for flow in a pi<strong>pe</strong>.


Solution in SI Units<br />

Using Equation 2-3, the shear stress at the wall is<br />

0.053(3000)<br />

�w = �� = 22 Pa<br />

4 × 1.8<br />

since the inner radius of the pi<strong>pe</strong> is 26.5 mm.<br />

At a radius of 20 mm the shear stress is<br />

r 20<br />

� = �w� � = 22� �<br />

At a radius of 12 mm the shear stress is<br />

r 12<br />

� = �w� � = 22� �<br />

At the center of the pi<strong>pe</strong> the shear stress is<br />

r 0<br />

� = �w� � = 22� �<br />

= 16.6 Pa<br />

= 10 Pa<br />

� = 0 Pa<br />

26.5<br />

Solution in USCS Units<br />

Using Equation 2-3, the shear stress at the wall is<br />

2.086 in (0.435 psi)<br />

�w = ��� = 0.0032 psi<br />

4 × 5.9 × 12<br />

since the inner radius of the pi<strong>pe</strong> is 1.043 in.<br />

At a radius of 20 mm (0.787 in) the shear stress is<br />

r<br />

0.787<br />

� = �w� � = 0.0032� �<br />

� = 0.0024 psi<br />

1.043<br />

At a radius of 12 mm (0.472 in) the shear stress is<br />

0.472<br />

� = 0.0032��� = 0.00145 psi<br />

1.043<br />

At the center of the pi<strong>pe</strong> � = 0<br />

FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

2-2 REYNOLDS NUMBER AND<br />

FLOW REGIMES<br />

� Ri<br />

� Ri<br />

� Ri<br />

� Ri<br />

� 26.5<br />

� 26.5<br />

Determining the magnitude of friction was historically a controversial topic until the end<br />

of the 19th century. The great disagreement was between the practical engineers and the<br />

theory of hydrodynamics.<br />

Hager (in 1839) and Poiseville (in 1840) demonstrated that under certain conditions<br />

friction was a linear function of the s<strong>pe</strong>ed of flow. In 1858, Darcy demonstrated that under<br />

other conditions friction was in fact proportional to the square of the mean s<strong>pe</strong>ed of the<br />

flow. By 1883, Reynolds had demonstrated that both Poiseville and Darcy were correct, as<br />

the mechanics of flows were fundamentally different at very low s<strong>pe</strong>eds and at high s<strong>pe</strong>eds.<br />

2.3


2.4 CHAPTER TWO<br />

Through nondimensional analysis, Reynolds demonstrated that under certain fixed<br />

conditions, the transition from a laminar Poiseville flow to a turbulent Darcy flow was<br />

based on the ratio of the inertia forces to the viscous forces. In his honor, such a ratio is<br />

now called the Reynolds number:<br />

Re = = = (2-4)<br />

where<br />

� = density of the fluid<br />

� = absolute or “dynamic” viscosity<br />

� = kinematic viscosity (defined as the absolute viscosity divided by the density of the<br />

liquid)<br />

U = average velocity of the flow<br />

The kinematic viscosity is the absolute (or dynamic viscosity) divided by the density.<br />

Its unit of measurement are, strictly s<strong>pe</strong>aking, m2 /s or ft2 Inertia forces �UDi UDi �� � �<br />

Viscous forces � �<br />

/sec. Another unit used is the<br />

centistokes, which is obtained by dividing the absolute viscosity in centipoises by the s<strong>pe</strong>cific<br />

gravity of the fluid. A centipoise is equivalent to one milli-Pascal-second.<br />

One unit for kinematic viscosity used in the oil industry (but of limit use in the mining<br />

industry) is the seconds Saybolt universal (or SSU). For values of kinematic viscosity<br />

larger than 70 centistokes (cst), the following formula is recommended by the Hydraulic<br />

Institute (1990):<br />

SSU = centistokes × 4.635<br />

In simplified terms, it may be said that two geometrically similar bodies immersed in a<br />

fluid will develop inertia and viscous forces in a constant ratio when body forces are negligible.<br />

Since Reynolds develo<strong>pe</strong>d his theory, his approach has been has been extended to other<br />

fluids. Modern aerodynamics uses the chord of the wing aerofoil instead of the pi<strong>pe</strong> diameter<br />

as the distance parameter for Equation 2-4. In Chapter 3, the concept of the particle<br />

Reynolds number based on a characteristic particle diameter shall be introduced.<br />

Figure 2.2 presents the equations of the Reynolds number for different sha<strong>pe</strong>s and flows.<br />

For pi<strong>pe</strong> flows, the critical Reynolds number is considered to be between 2300 and<br />

2800. In many pi<strong>pe</strong>s, flow becomes unstable above a Reynolds number of 2300 and slides<br />

into a transition regime before converting into turbulent motion.<br />

2-3 FRICTION FACTORS<br />

The Fanning friction factor is a nondimensional number defined as the ratio of the wall<br />

shear stress to the dynamic pressure of the flow:<br />

fN = � (2-5)<br />

2 �U /2<br />

Users of USCS units should use slugs/ft3 for density and not the more commonly used<br />

units of pounds <strong>pe</strong>r cubic feet. The conversion between these two is gc or 32.2 ft/sec2 .<br />

Substituting Equation 2-3 into Equation 2-5<br />

�U 2<br />

�PDI � 4L<br />

� w<br />

� 2<br />

f N = � � /� � (2-6)


c<br />

Airfoil chord "c"<br />

Re = U c /<br />

Annular flow<br />

Re= ( U/ )*2 r + r - r<br />

2 2 2<br />

0 i m<br />

2 2<br />

0 i 10 0 I<br />

where r = (r - r )/2.3 log (r /r )<br />

m<br />

FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

Close parallel plates<br />

Equation 2-6 clearly indicates that the friction factor is de<strong>pe</strong>ndent on the flow. It will<br />

be demonstrated that there is a relationship between the friction factor and the Reynolds<br />

number of the flow. In USCS units, density is expressed in slugs/ft 3 .<br />

Example 2-2<br />

The <strong>slurry</strong> in Example 2-1 has a s<strong>pe</strong>cific gravity of 1.2 , or a density of 1200 kg/m3 . If the<br />

s<strong>pe</strong>ed of the flow is 2 m/s (6.56 ft/s), determine the Fanning friction factor.<br />

Solution in SI Units<br />

Using the shear stress calculated in Example 2-1 as 22 Pa and inserting it into Equation 2-<br />

5, the Fanning friction factor can be calculated as<br />

22<br />

fN = �� = 0.0092<br />

2 1200 × 2 × 0.5<br />

Solutions in USCS Units<br />

Assume the density of water to be 62.3 lbm/ft3 . Since s<strong>pe</strong>cific gravity equals 1.2, the density<br />

of the <strong>slurry</strong> is 1.2 × 62.3 = 74.76 lbs/ft3 . To convert lbs/ft3 , into slugs/ft3 complete<br />

the following equation:<br />

= 2.32 slugs/ft3 74.76<br />

�<br />

32.2<br />

In Example 2-1, the wall shear stress was determined to be 0.0032 psi. Using equation<br />

2.5<br />

0.0032 × 144<br />

fN = �� = 0.0092<br />

2 2.32 × 6.56 × 0.5<br />

h<br />

2.5<br />

Use inner diameter for sphere<br />

for pi<strong>pe</strong> flow use diameter<br />

Re= U D /<br />

i<br />

Re = [U h/2 ] 32/3<br />

FIGURE 2-2 Definition of Reynolds number for various sha<strong>pe</strong>s.<br />

D I<br />

Re= U d /<br />

p<br />

d p


2.6 CHAPTER TWO<br />

2-3-1 Laminar Friction Factors<br />

The friction or resistance of a body to motion is confined to a viscous layer at the wall or<br />

surface of the body. This layer is called the boundary layer. Understanding this phenomenon<br />

remained illusive until the end of the 19th century. At the turn of the 20th century,<br />

Prandtl develo<strong>pe</strong>d the principles of boundary layer theory.<br />

To understand the basic principles of the boundary layer, let us start by examining the<br />

flow between closely related plates at low s<strong>pe</strong>ed. With one plate stationary and the other<br />

moving at a s<strong>pe</strong>ed U, a linear velocity profile is established.<br />

The wall shear stress �w is established as<br />

�U<br />

�w = �<br />

h<br />

by Newton’s law, where h is the spacing between plates and<br />

U<br />

u = y� �<br />

� (2-7)<br />

h<br />

With y as the vertical ordinate from the stationary plate, the velocity gradient is defined as<br />

du U<br />

� = � = rate of shearing strain or shear rate (2-8)<br />

d� H<br />

Thus, the dynamic viscosity is<br />

� Shear Stress<br />

� = � = ���<br />

(2-9)<br />

du/d� Rate of Shear Strain<br />

Equation 2-8 is the basis for boundary layer theory. Instead of a moving plate at a velocity<br />

U, the velocity of the flow outside the boundary layer is studied (see Figure 2-3).<br />

The shear rate is not necessarily linear. For example flow around an aircraft airfoil can<br />

be attached, stagnant at a point, or can even reverse flow after separation.<br />

Laminar flow in a pi<strong>pe</strong> is described by the Hagen–Poiseville equation:<br />

�P 32�U<br />

� = � (2-10)<br />

2 L D i<br />

Up<strong>pe</strong>r plate moves at s<strong>pe</strong>ed U<br />

h u=U<br />

y<br />

y<br />

h w<br />

FIGURE 2-3 Linear velocity distribution due to a plate moving at a s<strong>pe</strong>ed U above a stationary<br />

plate.


which can be rearranged as<br />

FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

� = � �� � = 2 �P D i Di�P Di � � ������ (2-11)<br />

L 32U 4L 8U<br />

� = � (2-12)<br />

8U/Di<br />

or the ratio of the wall shear stress and the mean velocity gradient.<br />

The term (8U/Di) is often called the pi<strong>pe</strong> (or pi<strong>pe</strong>line) shear rate in laminar flow. From<br />

Equation 2-12<br />

�8U<br />

�w = (2-13a)<br />

� Di<br />

Substituting Equation 2-12 into Equation 2-5, the Fanning friction factor for a laminar<br />

flow can be expressed as<br />

�U<br />

fN = � �/� � (2-14)<br />

2<br />

�8U<br />

� �<br />

Di 2<br />

16� 16<br />

fN = � = � (2-15)<br />

�VDi Re<br />

The Fanning friction factor is more commonly found in reference publications on<br />

chemical engineering. Another friction factor used by mechanical engineers is the Darcy<br />

friction factor:<br />

f Darcy = 4 × f Fanning<br />

To avoid confusion, in the book the symbols f D will be used for Darcy friction factor<br />

and f N will be used for Fanning friction factor.<br />

Flow in a laminar regime is considered to be inde<strong>pe</strong>ndent of pi<strong>pe</strong> roughness.<br />

Example 2-3<br />

A viscous fluid is flowing in a laminar flow at a s<strong>pe</strong>ed of 1m/s (or 3.28 ft/s) in a pi<strong>pe</strong><br />

with an inner diameter of 336.6 mm (13.25 in). The measured pressure drop over a distance<br />

of 200 m (656 ft) is 8400 Pa (1.22 psi). The density of the fluid is 855 kg/m 3 (SG<br />

= 0.855).<br />

Determine an equivalent viscosity for the pi<strong>pe</strong>line fluid, the Reynolds number, and the<br />

friction factor.<br />

Solution in SI Units<br />

From Equation 2-3, the shear stress at the wall is<br />

0.3366 × 8400<br />

�w = �� = 3.53 Pa<br />

4 × 200<br />

The Fanning Friction Factor from Equation 2.5 is<br />

2�w 2 × 3.53<br />

fN = � = � = 0.0083<br />

2<br />

2<br />

�V 855 × 1<br />

� w<br />

2.7


2.8 CHAPTER TWO<br />

The equivalent pi<strong>pe</strong>line viscosity from Equation 2-12 is<br />

�w 3.53 × 0.3366<br />

� = � = �� = 0.148 Pa-s<br />

8U/D 8 × 1<br />

The Reynolds Number from Equation 2-4<br />

�UD 855 × 1 × 0.3366<br />

Re = � = �� = 1938<br />

� 0.148<br />

Check on friction factor:<br />

16<br />

fN = � = 0.00823<br />

1938<br />

Solution in USCS Units<br />

density = = 1.654 slugs/ft3 Since it was stated that the pressure drop = 1.215 psi, over a distance of 656.2 ft the<br />

shear stress is:<br />

�w = = 0.0005 psi<br />

The Fanning Friction Factor from Equation 2.5 is<br />

fN = = = 0.0081<br />

The equivalent pi<strong>pe</strong>line viscosity converting the shear stress from psi to lbf/ft2 , 0.0005<br />

× 144 = 0.072 lbf/ft2 , is<br />

� = = = 0.003 lbf-sec/ft2 0.8555 × 62.3<br />

��<br />

32.2<br />

13.25 × 1.2<br />

��<br />

4 × 656.2 × 12<br />

2�w 2 × 0.0005 × 144<br />

� ��<br />

2<br />

2<br />

�V 1.654 × 3.28<br />

Di�w (13.25/12) × 0.072<br />

� ��<br />

8V 8 × 3.28<br />

The Reynolds number is<br />

1.654 × 3.28 × 13.25/12<br />

Re = ��� = 1997<br />

0.003<br />

Check on friction factor:<br />

16<br />

fN = � = 0.008<br />

1997<br />

2-3-2 Transition Flow Friction Factor<br />

The transition from a laminar to a turbulent flow is difficult to describe.<br />

For Reynolds numbers up to 3 × 106 , Wasp et al. (1977) recommended the use of Nikuradse<br />

equation for the Fanning coefficient:<br />

0.0553<br />

fN = 0.0008 + � (2-16)<br />

0.237 Re


FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

2-3-3 Friction Factor in Turbulent Flow<br />

In turbulent flow, the roughness of the pi<strong>pe</strong> becomes an important factor in determining<br />

the friction factor. The roughness � is measured in units of length. Up to a certain limiting<br />

Reynolds number Re, the friction factor can be expressed by the following Colebrook<br />

equation:<br />

= 4 loge� � + 3.48 – 4 loge�1 + 9.35 � (2-17)<br />

A more simplified equation for the Darcy factor is called the Prandtl–Colebrook equation:<br />

= –2 log10� + � (2-18)<br />

A popular equation used by mechanical engineers (Lindeburg, 1997) as it is explicit<br />

and does not require tedious iterations is the Swamee–Jain equation. It is suitable for the<br />

range of Reynolds numbers between 5000 and 100,000,000:<br />

0.25<br />

fD = ���<br />

(2-19)<br />

��� log10��� / D<br />

� + (5.74/Re<br />

3.<br />

7<br />

0.9 )�� 2<br />

1<br />

Di Di � � ��<br />

�fN� 2�<br />

2� Re�fN�<br />

1<br />

2.51 �<br />

� � �<br />

�fD�<br />

Re�fN� 3.7 Di<br />

Because the <strong>slurry</strong> flows occur at Reynolds Numbers smaller than 100,000,000, Equation<br />

2.19 is satisfactory in the context of this <strong>handbook</strong>.<br />

Example 2-4<br />

Using the Swamee–Jain equation, determine the friction factor for a flow of 3500 US gpm<br />

in an 18� OD pi<strong>pe</strong> with a wall thickness of 0.375 in. The fluid has a s<strong>pe</strong>cific gravity of<br />

1.02 and a dynamic viscosity of 2.7 × 10 –5 lbf-sec/ft 2 .<br />

Solutions in SI Units (For conversion factors refer to the Ap<strong>pe</strong>ndix at the end of this<br />

book.)<br />

Q = = 0.221 m3 3500 × 3.785<br />

�� /s<br />

60000<br />

ID of pi<strong>pe</strong> = (18 – 2 × 0.375) = 17.25 in (0.438 m)<br />

Area of flow = � × 0.25 × 0.438 2 = 0.1506 m 2<br />

0.221<br />

Average velocity of flow = � = 1.467m/s<br />

0.1506<br />

Density = 1.02 × 1000 = 1020 kg/m 3<br />

� = 2.7 × 10 –5 × 47.88 = 0.00129 Pa.s<br />

1020 × 1.467 × 0.438<br />

Re = ��� = 506975<br />

0.00129<br />

2.9


2.10 CHAPTER TWO<br />

The absolute roughness � of steel pi<strong>pe</strong>s is 6 × 10 –5 m. Relative roughness is<br />

6 × 10<br />

�/Di = = 0.000137<br />

–5<br />

�<br />

0.438<br />

fD = = 0.01486<br />

Solutions in USCS Units<br />

Q = 3500 × 0.1337 = 467.95 ft3 0.25<br />

�����<br />

[log10(0.000137/3.7 + 5.74/506975<br />

/min<br />

0.9 )] 2<br />

ID of pi<strong>pe</strong> = (18 – 2 × 0.375) = 17.25 in or 1.4375 ft<br />

Area of flow = � × 0.25 × 1.4375 2 = 1.623 ft 2<br />

467.95<br />

Velocity of flow = � = 255.3 ft/min or 4.80 ft/s<br />

1.623<br />

Density = = 1.973 slugs/ft3 1.02 × 62.3<br />

��<br />

32.2<br />

1.973 × 4.80 × 1.4375<br />

Re = ��� = 504,779<br />

2.7 × 10<br />

The absolute roughness of steel is 0.0002 ft. Relative roughness is<br />

–5<br />

0.0002<br />

� 1.4375<br />

= 0.000139<br />

fD = 0.25/[log10(0.000139/3.7 + 5.74/5047790.9 )] 2 = 0.0149<br />

The Colebrook and Prandtl–Colebrook formulas are limited to a certain range of<br />

Reynolds number magnitude. At high Reynolds numbers, the friction factor becomes inde<strong>pe</strong>ndent<br />

of the Reynolds number. The value of the Reynolds number beyond which the<br />

friction factor is inde<strong>pe</strong>ndent is calculated using the following equation:<br />

Re = 70 �� = 70 Di 2 Di 8<br />

� � � �<br />

� fN � ��fD<br />

The region for which equation 2-19 applies is shown on the Moody diagram to be to<br />

the right of the the dashed curve (Figure 2-4).<br />

The equations established so far have been develo<strong>pe</strong>d for clear water and do not apply<br />

for plastic fluids or liquids carrying coarse particles. They can apply for any other singlephase<br />

Newtonian liquids. (These terms will be explained in Chapter 3).<br />

Most fluid dynamics books publish data for linear roughness based on Moody’s work.<br />

Such values are applicable to water. However, tests conduced on <strong>slurry</strong> pi<strong>pe</strong>lines can<br />

yield different values, due to the erosion of pi<strong>pe</strong>, wear and tear of rubber linings, etc.<br />

The Moody diagram is a general graph for the Darcy factor versus the Reynolds number.<br />

It is applicable to a very large number of different pi<strong>pe</strong> materials. There are four principal<br />

pi<strong>pe</strong> materials associated with <strong>slurry</strong> flows: plastic pi<strong>pe</strong>s [high-density polyethylene<br />

(HDPE)], plain steel pi<strong>pe</strong>s, rubber-lined steel pi<strong>pe</strong>s, and concrete pi<strong>pe</strong>s. The absolute


2.11<br />

FIGURE 2-4 Moody diagram for the friction factor versus the Reynolds number for pi<strong>pe</strong> flow (Reproduced from V. L. Streeter, Fluid<br />

Mechanics, McGraw-Hill, 1971. Reproduced by <strong>pe</strong>rmission of McGraw-Hill, Inc.)


2.12 CHAPTER TWO<br />

roughness of these materials is presented in Table 2-1. Plain steel pi<strong>pe</strong>s and rubber-lined<br />

steel pi<strong>pe</strong>s are the most common, but HDPE and HDPE-lined steel pi<strong>pe</strong>s have gained in<br />

importance in the last quarter of the 20th century. One of the concentrate pi<strong>pe</strong>lines used in<br />

Escondida, Chile featured a long section of gravity flow in HDPE pi<strong>pe</strong>.<br />

Certain reference books show a roughness of steel of 0.045–0.05 mm. This is difficult<br />

to maintain in steel pi<strong>pe</strong>s carrying slurries as they are often subject to erosion and corrosion.<br />

For this reason, the author recommends the use of a slightly higher roughness of the<br />

order of 0.06 mm in friction calculations.<br />

The dimensions of plain steel pi<strong>pe</strong>s, their pressure ratings, and relative roughness are<br />

presented in Table 2-2. It is obvious that steel pi<strong>pe</strong>s are limited in pressure rating to<br />

3000 psi. This criterion is essential when considering location of booster pump stations<br />

or chokes. In the case of <strong>slurry</strong> pi<strong>pe</strong>lines, the thickness of the pi<strong>pe</strong>s is selected on the<br />

basis of<br />

� Pressure<br />

� Corrosion allowance<br />

� Wear allowance<br />

Because of the wear allowance, erosion, and corrosion, the rating of <strong>slurry</strong> pi<strong>pe</strong>s is<br />

lower than presented in Table 2-2. Steel pi<strong>pe</strong>s may be hardened for carrying coarse <strong>slurry</strong><br />

particles (larger than 6.5 mm or 1 – 4�), or sacrificial thickness is used to resist abrasion.<br />

The pressure rating as presented in Table 2-2 should be used as a starting point for the<br />

design calculations, and the appropriate allowance should be made for wear. Reclaimed<br />

water pi<strong>pe</strong>lines use the pressure ratings as in Table 2-2. The roughness and inner diameter<br />

of reclaimed water pi<strong>pe</strong>s may change due to scaling and deposition of lime.<br />

Steel pi<strong>pe</strong>s are rubber lined to a typical thickness of 6 mm (or 0.25�) for small sizes of<br />

pi<strong>pe</strong>s [< 150 mm (6�), 9.5 mm ( 3 – 8�)], and 13 mm ( 1 – 2�) for pi<strong>pe</strong> sizes up to 24�. Larger pi<strong>pe</strong>s<br />

may be custom lined. Lining is done in an autoclave and the rubber is cured under steam.<br />

Rubber lining is limited to pumping coarse material up to a size of 6 mm (� 1 – 4�). Rubber<br />

does not contribute to the pressure rating of steel pi<strong>pe</strong>s. Table 2-3 presents the dimensions<br />

and relative roughness of rubber-lined steel pi<strong>pe</strong>s.<br />

The dimensions of plain HDPE pi<strong>pe</strong>s (not HDPE-lined steel) for pressures up to 110<br />

psi (760 kPa) are listed in Table 2-4. The dimensions of HDPE pi<strong>pe</strong>s for pressures in the<br />

range of 125 to 300 psi (863–2070 kPa) are presented in Table 2-5. These dimensions are<br />

slightly different than metric pi<strong>pe</strong>s. HDPE is not a magic material but can withstand the<br />

abrasion of taconite and some coarse laterites. As with rubber, there must be a cut-off size<br />

of particle size beyond which the use of HDPE is not acceptable. Very little has been published<br />

on this subject.<br />

The use of concrete pi<strong>pe</strong>s is often associated with gravity flows.<br />

TABLE 2-1 Absolute Roughness of New Materials Used in Slurry Pi<strong>pe</strong>s<br />

Description Roughness (m) Roughness (ft)<br />

Plastic pi<strong>pe</strong>s, PVC, ABS, HDPE 1.5 × 10 –6 0.000004921<br />

Steel pi<strong>pe</strong>s 6.0 × 10 –5 0.000197<br />

Rubber-lined pi<strong>pe</strong>s 0.00015 0.000492<br />

Concrete pi<strong>pe</strong>s 0.0012 0.00394


FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

TABLE 2-2 Size, Rating, and Relative Roughness of Plain Steel Pi<strong>pe</strong>s to<br />

U.S. Dimensions*<br />

2.13<br />

Pi<strong>pe</strong> Allowed Relative<br />

Outside Wall internal working roughness<br />

Denomination diameter thickness diameter pressure (psi) for new<br />

(inch) (inch) (inch) (inch) to 650 °F pi<strong>pe</strong><br />

2� Sch 40 2.375 0.154 2.067 1159 0.001143<br />

2� Sch 80 0.218 1.939 2038 0.001212<br />

2� Sch 160 0.344 1.687 3890 0.001400<br />

XX 0.436 1.503 5356 0.001572<br />

3� Sch 40 3.500 0.216 3.068 1341 0.000770<br />

3� Sch 80 0.300 2.900 2129 0.000815<br />

3� Sch 160 0.438 2.624 3495 0.000900<br />

XX 0.600 2.300 5252 0.001027<br />

4� Sch 40 4.500 0.237 4.026 1191 0.000587<br />

4� Sch 80 0.337 3.826 1905 0.000617<br />

4� Sch 120 0.438 3.624 2663 0.000652<br />

4� Sch 160 0.531 3.438 3387 0.000687<br />

XX 0.674 3.152 4553 0.000749<br />

5� Sch 40 5.633 0.257 5.117 1071 0.000461<br />

5� Sch 80 0.375 4.883 1950 0.000484<br />

5� Sch 120 0.500 4.633 2502 0.000510<br />

5� Sch 160 0.625 4.383 3284 0.000539<br />

XX 0.750 4.133 4098 0.000572<br />

6� Sch 40 6.625 0.280 6.065 1000 0.000389<br />

6� Sch 80 0.432 5.761 1739 0.000410<br />

6� Sch 120 0.562 5.501 2394 0.000429<br />

6� Sch 160 0.719 5.187 3215 0.000455<br />

XX 0.864 4.897 4004 0.000482<br />

8� Sch 20 8.625 0.250 8.125 655 0.000291<br />

8� Sch 30 0.277 8.071 752 0.000293<br />

8� Sch 40 0.322 7.981 916 0.000296<br />

8� Sch 60 0.406 7.813 1225 0.000302<br />

8� Sch 80 0.500 7.625 1577 0.000310<br />

8� Sch 100 0.594 7.437 1935 0.000318<br />

8� Sch 120 0.719 7.187 2422 0.000329<br />

8� Sch 140 0.812 7.001 2792 0.000337<br />

XX 0.875 6.875 3046 0.000344<br />

8� Sch 160 0.906 6.813 3173 0.000347<br />

10� Sch 20 10.75 0.250 10.250 523 0.000230<br />

10� Sch 30 0.307 10.136 688 0.000233<br />

10� Sch 40 S 0.365 10.020 856 0.000236<br />

10� Sch 60 X 0.500 9.750 1255 0.000243<br />

10� Sch 80 0.594 9.562 1537 0.000247<br />

10� Sch 100 0.719 9.312 1918 0.000254<br />

10� Sch 120 0.844 9.062 2308 0.000261<br />

10� Sch 140 XX 1.000 8.750 2804 0.000270<br />

10� Sch 160 1.125 8.500 3211 0.000278<br />

(continued)


2.14 CHAPTER TWO<br />

TABLE 2-2 Continued<br />

Pi<strong>pe</strong> Allowed Relative<br />

Outside Wall internal working roughness<br />

Denomination diameter thickness diameter pressure (psi) for new<br />

(inch) (inch) (inch) (inch) to 650 °F pi<strong>pe</strong><br />

12� S 12.750 0.250 12.250 440 0.000193<br />

12� Sch 40 0.330 12.090 634 0.000195<br />

12� X 0.375 12.000 744 0.000197<br />

12� Sch 60 X 0.406 11.938 820 0.000198<br />

12� Sch 80 0.500 11.750 1052 0.000201<br />

12� Sch 100 0.562 11.626 1207 0.000203<br />

12� Sch 120 XX 0.688 11.374 1526 0.000208<br />

12� Sch 140 0.844 11.062 1927 0.000214<br />

12� Sch 160 1.00 10.750 2337 0.000220<br />

1.125 10.500 2672 0.000225<br />

1.312 10.126 3183 0.000233<br />

14� Sch 10 14.000 0.250 13.500 401 0.000175<br />

14� Sch 20 0.330 13.376 537 0.000177<br />

14� Sch 30 S 0.375 13.250 676 0.000178<br />

14� Sch 40 0.400 13.124 817 0.00180<br />

14� X 0.500 13.000 956 0.000182<br />

14� Sch 60 0.594 12.812 1169 0.000184<br />

14� Sch 80 0.750 12.500 1528 0.000189<br />

14� Sch 100 0.938 12.124 1969 0.000195<br />

14� Sch 120 1.062 11.876 2265 0.000199<br />

14� Sch 140 1.250 11.500 2724 0.000205<br />

14� Sch 160 1.406 11.188 3112 0.000211<br />

16� Sch 10 16.000 0.250 15.500 350 0.000152<br />

16� Sch 20 0.312 15.376 469 0.000154<br />

16� Sch 30 S 0.375 15.250 590 0.000155<br />

16� Sch 40 X 0.500 15.000 834 0.000157<br />

16� Sch 60 0.656 14.688 1142 0.000161<br />

16� Sch 80 0.844 14.312 1520 0.000165<br />

16� Sch 100 1.031 13.938 1903 0.000169<br />

16� Sch 120XX 1.219 13.562 2296 0.000176<br />

16� Sch 140 1.438 13.124 2764 0.000180<br />

16� Sch 160 1.594 12.812 3104 0.000184<br />

18� Sch 10 18.000 0.250 17.500 312 0.000135<br />

18� Sch 20 0.312 17.376 416 0.000136<br />

18� S 0.375 17.250 524 0.000137<br />

18� Sch 30 0.438 17.124 632 0.000138<br />

18� X 0.500 17.000 739 0.000139<br />

18� Sch 40 0.562 16.876 847 0.000140<br />

18� Sch 60 0.750 16.500 1178 0.000143<br />

18� Sch 80 0.938 16.124 1514 0.000147<br />

18� Sch 100 1.156 15.688 1911 0.000151<br />

18� Sch 120 1.375 15.225 2318 0.000155<br />

18� Sch 140 1.562 14.876 2673 0.000159<br />

18� Sch 160 1.781 14.438 3096 0.000164


TABLE 2-2 Continued<br />

FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

2.15<br />

Pi<strong>pe</strong> Allowed Relative<br />

Outside Wall internal working roughness<br />

Denomination diameter thickness diameter pressure (psi) for new<br />

(inch) (inch) (inch) (inch) to 650 °F pi<strong>pe</strong><br />

20� Sch 10 20.000 0.250 19.500 280 0.000121<br />

20� Sch 20 S 0.375 19.250 471 0.000123<br />

20� Sch 30 X 0.500 19.000 664 0.000124<br />

20� Sch 40 0.594 18.812 811 0.000126<br />

20� Sch 60 0.812 18.376 1155 0.000129<br />

20� Sch 80 1.031 17.938 1507 0.000132<br />

20� Sch 100 1.281 17.438 1917 0.000135<br />

20� Sch 120 1.500 17.000 2284 0.000139<br />

20� Sch 140 1.750 16.500 2710 0.000143<br />

20� Sch 160 1.969 16.082 3091 0.000147<br />

24� Sch 10 24.000 0.250 23.500 233 0.0001005<br />

24� Sch 20 S 0.375 23.250 392 0.0001016<br />

24� X 0.500 23.000 552 0.0001027<br />

24� Sch 30 0.562 22.876 635 0.000103<br />

24� Sch 40 0.688 22.624 795 0.000104<br />

24� Sch 60 0.969 22.062 1165 0.000107<br />

24� Sch 80 1.219 21.562 1500 0.000109<br />

24� Sch 100 1.513 20.938 1927 0.000113<br />

24� Sch 120 1.812 20.376 2319 0.000116<br />

24� Sch 140 2.062 19.875 2674 0.000119<br />

24� Sch 160 2.344 19.312 3083 0.000122<br />

30� Sch 10 30.000 0.312 29.376 254 0.0000804<br />

30� S 0.375 29.250 313 0.0000807<br />

30� Sch 20 X 0.500 29.00 440 0.0000814<br />

30� Sch 30 0.625 28.750 568 0.0000822<br />

36� Sch 10 36.000 0.312 35.376 207 0.0000668<br />

36� S 0.375 35.250 260 0.0000670<br />

36� Sch 20 X 0.500 35.000 366 0.0000675<br />

36� Sch 30 0.625 34.750 473 0.0000679<br />

36� Sch 40 0.750 34.500 580 0.0000685<br />

42 S 42.000 0.375 41.250 223 0.0000573<br />

0.500 41.000 313 0.0000576<br />

48 S 48.000 0.375 47.250 195 0.0000499<br />

0.500 47.000 274 0.0000503<br />

*Dimensions are based on ANSI B36.1 and B31.1. Absolute roughness of steel is taken as 60 �m.


2.16 CHAPTER TWO<br />

TABLE 2-3 Size and Relative Roughness of Rubber-Lined Steel Pi<strong>pe</strong>s to U.S.<br />

Dimensions*<br />

Relative<br />

Outside Wall Rubber Pi<strong>pe</strong> internal roughness for<br />

Denomination diameter thickness thickness diameter rubber-lined<br />

(inch) (inch) (inch) (inch) (inch) pi<strong>pe</strong><br />

2� Sch 40<br />

2� Sch 80<br />

2� Sch 160<br />

XX<br />

3� Sch 40<br />

3� Sch 80<br />

3� Sch 160<br />

XX<br />

4� Sch 40<br />

4� Sch 80<br />

4� Sch 120<br />

4� Sch 160<br />

XX<br />

5� Sch 40<br />

5� Sch 80<br />

5� Sch 120<br />

5� Sch 160<br />

XX<br />

6� Sch 40<br />

6� Sch 80<br />

6� Sch 120<br />

6� Sch 160<br />

XX<br />

8� Sch 20<br />

8� Sch 30<br />

8� Sch 40<br />

8� Sch 60<br />

8� Sch 80<br />

8� Sch 100<br />

8� Sch 120<br />

8� Sch 140<br />

XX<br />

8� Sch 160<br />

10� Sch 20<br />

10� Sch 30<br />

10� Sch 40 S<br />

10� Sch 60 X<br />

10� Sch 80<br />

10� Sch 100<br />

10� Sch 120<br />

10� Sch 140 XX<br />

10� Sch 16<br />

2.375<br />

3.500<br />

4.500<br />

5.633<br />

6.625<br />

8.625<br />

10.75<br />

0.154<br />

0.218<br />

0.344<br />

0.436<br />

0.216<br />

0.300<br />

0.438<br />

0.600<br />

0.237<br />

0.337<br />

0.438<br />

0.531<br />

0.674<br />

0.257<br />

0.375<br />

0.500<br />

0.625<br />

0.750<br />

0.280<br />

0.432<br />

0.562<br />

0.719<br />

0.864<br />

0.250<br />

0.277<br />

0.322<br />

0.406<br />

0.500<br />

0.594<br />

0.719<br />

0.812<br />

0.875<br />

0.906<br />

0.250<br />

0.307<br />

0.365<br />

0.500<br />

0.594<br />

0.719<br />

0.844<br />

1.000<br />

1.125<br />

0.25<br />

0.25<br />

0.25<br />

0.25<br />

0.25<br />

0.375<br />

0.375<br />

1.559<br />

1.431<br />

1.179<br />

0.995<br />

2.560<br />

2.392<br />

2.116<br />

1.792<br />

3.518<br />

3.138<br />

3.116<br />

2.930<br />

2.664<br />

4.609<br />

4.375<br />

4.125<br />

3.875<br />

3.625<br />

5.557<br />

5.253<br />

4.993<br />

4.679<br />

4.389<br />

7.375<br />

7.732<br />

7.231<br />

7.063<br />

6.875<br />

6.687<br />

6.437<br />

6.251<br />

6.125<br />

6.063<br />

9.500<br />

9.386<br />

9.27<br />

9.00<br />

8.812<br />

8.562<br />

8.312<br />

8.000<br />

7.750<br />

0.003788<br />

0.004127<br />

0.005009<br />

0.005935<br />

0.002307<br />

0.002469<br />

0.002790<br />

0.003295<br />

0.001679<br />

0.00178<br />

0.001895<br />

0.002015<br />

0.002233<br />

0.001281<br />

0.00135<br />

0.001431<br />

0.001524<br />

0.001629<br />

0.001063<br />

0.001124<br />

0.001183<br />

0.001262<br />

0.001345<br />

0.000801<br />

0.000807<br />

0.000817<br />

0.000836<br />

0.000859<br />

0.000883<br />

0.000917<br />

0.000945<br />

0.000964<br />

0.000974<br />

0.000621<br />

0.000629<br />

0.000637<br />

0.000656<br />

0.000670<br />

0.000689<br />

0.000710<br />

0.000738<br />

0.000762


TABLE 2-3 Continued<br />

FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

2.17<br />

Relative<br />

Outside Wall Rubber Pi<strong>pe</strong> internal roughness for<br />

Denomination diameter thickness thickness diameter rubber-lined<br />

(inch) (inch) (inch) (inch) (inch) pi<strong>pe</strong><br />

12� S<br />

12� Sch 40<br />

12� X<br />

12� Sch 60 X<br />

12� Sch 80<br />

12� Sch 100<br />

12� Sch 120 XX<br />

12� Sch 140<br />

12� Sch 160<br />

14� Sch 10<br />

14� Sch 20<br />

14� Sch 30 S<br />

14� Sch 40<br />

14� X<br />

14� Sch 60<br />

14� Sch 80<br />

14� Sch 100<br />

14� Sch 120<br />

14� Sch 140<br />

14� Sch 160<br />

16� Sch 10<br />

16� Sch 20<br />

16� Sch 30 S<br />

16� Sch 40 X<br />

16� Sch 60<br />

16� Sch 80<br />

16� Sch 100<br />

16� Sch 120XX<br />

16� Sch 140<br />

16� Sch 160<br />

18� Sch 10<br />

18� Sch 20<br />

18� S<br />

18� Sch 30<br />

18� X<br />

18� Sch 40<br />

18� Sch 60<br />

18� Sch 80<br />

18� Sch 100<br />

18� Sch 120<br />

18� Sch 140<br />

18� Sch 160<br />

12.750<br />

14.000<br />

16.000<br />

18.000<br />

0.250<br />

0.330<br />

0.375<br />

0.406<br />

0.500<br />

0.562<br />

0.688<br />

0.844<br />

1.00<br />

1.125<br />

1.312<br />

0.250<br />

0.330<br />

0.375<br />

0.400<br />

0.500<br />

0.594<br />

0.750<br />

0.938<br />

1.062<br />

1.250<br />

1.406<br />

0.250<br />

0.312<br />

0.375<br />

0.500<br />

0.656<br />

0.844<br />

1.031<br />

1.219<br />

1.438<br />

1.594<br />

0.250<br />

0.312<br />

0.375<br />

0.438<br />

0.500<br />

0.562<br />

0.750<br />

0.938<br />

1.156<br />

1.375<br />

1.562<br />

1.781<br />

0.375<br />

0.375<br />

0.375<br />

0.375<br />

11.500<br />

11.340<br />

11.250<br />

11.188<br />

11.000<br />

10.870<br />

10.624<br />

10.312<br />

10.000<br />

9.750<br />

9.376<br />

12.750<br />

12.626<br />

12.500<br />

12.374<br />

12.250<br />

12.062<br />

11.750<br />

11.374<br />

11.126<br />

10.750<br />

10.438<br />

14.750<br />

14.626<br />

14.500<br />

14.250<br />

13.938<br />

13.562<br />

13.188<br />

12.668<br />

12.374<br />

12.062<br />

16.750<br />

16.626<br />

16.500<br />

16.374<br />

16.250<br />

16.126<br />

15.750<br />

15.374<br />

14.938<br />

14.500<br />

14.126<br />

13.688<br />

0.000513<br />

0.000521<br />

0.000525<br />

0.000528<br />

0.000537<br />

0.000543<br />

0.000556<br />

0.000572<br />

0.000591<br />

0.000606<br />

0.000629<br />

0.000463<br />

0.000468<br />

0.000472<br />

0.000477<br />

0.000482<br />

0.000489<br />

0.000503<br />

0.000519<br />

0.000531<br />

0.000549<br />

0.000566<br />

0.000400<br />

0.000404<br />

0.000407<br />

0.000414<br />

0.000424<br />

0.000435<br />

0.000448<br />

0.000466<br />

0.000477<br />

0.000489<br />

0.000353<br />

0.000355<br />

0.000358<br />

0.000361<br />

0.000363<br />

0.000366<br />

0.000375<br />

0.000384<br />

0.000395<br />

0.000407<br />

0.000418<br />

0.000431<br />

(continued)


2.18 CHAPTER TWO<br />

TABLE 2-3 Continued<br />

Relative<br />

Outside Wall Rubber Pi<strong>pe</strong> internal roughness for<br />

Denomination diameter thickness thickness diameter rubber-lined<br />

(inch) (inch) (inch) (inch) (inch) pi<strong>pe</strong><br />

20� Sch 10<br />

20� Sch 20 S<br />

20� Sch 30 X<br />

20� Sch 40<br />

20� Sch 60<br />

20� Sch 80<br />

20� Sch 100<br />

20� Sch 120<br />

20� Sch 140<br />

20� Sch 160<br />

24� Sch 10<br />

24� Sch 20 S<br />

24� X<br />

24� Sch 30<br />

24� Sch 40<br />

24� Sch 60<br />

24� Sch 80<br />

24� Sch 100<br />

24� Sch 120<br />

24� Sch 140<br />

24� Sch 160<br />

*Dimensions for steel are based on ANSI B36.1 and B31.1. Absolute roughness of rubber is taken<br />

as 150 �m.<br />

2-3-4 Hazen–Williams Formula<br />

In the United States, the Hazen–Williams formula is often used by civil engineers because<br />

it is inde<strong>pe</strong>ndent of the Reynolds number.<br />

S<strong>pe</strong>ed is calculated as:<br />

0.63 0.54 U = 1.319 CRH S in ft/s (2-20)<br />

S = Hv/L = ���<br />

(2-21)<br />

1.67 × C 1.85 1.17 × RH where Q gpm is expressed in US gallons <strong>pe</strong>r minute, and<br />

S = slo<strong>pe</strong> or head loss <strong>pe</strong>r unit length<br />

U = average velocity of fluid in ft/sec<br />

R H = hydraulic radius = area of pi<strong>pe</strong>/<strong>pe</strong>rimeter of pi<strong>pe</strong><br />

C = Surface roughness coefficient (refer to Table 2-6)<br />

In SI units:<br />

20.000<br />

24.000<br />

0.250<br />

0.375<br />

0.500<br />

0.594<br />

0.812<br />

1.031<br />

1.281<br />

1.500<br />

1.750<br />

1.969<br />

0.250<br />

0.375<br />

0.500<br />

0.562<br />

0.688<br />

0.969<br />

1.219<br />

1.513<br />

1.812<br />

2.062<br />

2.344<br />

0.375<br />

0.375<br />

1.85 Qgpm 18.750<br />

18.500<br />

18.250<br />

18.062<br />

17.626<br />

17.188<br />

16.688<br />

16.250<br />

15.750<br />

15.312<br />

22.750<br />

22.500<br />

22.250<br />

22.126<br />

21.874<br />

21.312<br />

20.812<br />

20.224<br />

19.626<br />

19.126<br />

18.562<br />

0.000315<br />

0.000319<br />

0.000324<br />

0.000327<br />

0.000335<br />

0.000344<br />

0.000354<br />

0.000363<br />

0.000375<br />

0.000386<br />

0.000259<br />

0.000260<br />

0.000263<br />

0.000265<br />

0.000267<br />

0.000270<br />

0.000277<br />

0.000284<br />

0.000292<br />

0.000309<br />

0.000318


FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

TABLE 2-4 Dimensions of North American HDPE Pi<strong>pe</strong>s for Pressure Ratings 50 psi<br />

to 110 psi at a Tem<strong>pe</strong>rature of 23°C (73.4 °F)<br />

2.19<br />

DR 32.5 DR 26 DR 21 DR 17 DR 15.5<br />

Pressure rating 50 psi 64 psi 80 psi 100 psi 110 psi<br />

Average Average Average Average Average Average<br />

outside inside inside inside inside inside<br />

Pi<strong>pe</strong> size diameter diameter diameter diameter diameter diameter<br />

(inch) (inch) (inch) (inch) (inch) (inch) (inch)<br />

3 3.500 3.214 3.147 3.063 3.021<br />

4 4.500 4.133 4.046 3.938 3.885<br />

5 5.563 5.109 5.001 4.870 4.802<br />

6 6.625 6.193 6.084 5.957 5.798 5.720<br />

7 7.125 6.661 6.544 6.406 6.237 6.150<br />

8 8.625 8.063 7.921 7.754 7.550 7.446<br />

10 10.750 10.048 9.874 9.665 9.410 9.279<br />

12 12.750 11.919 11.711 11.463 11.160 11.005<br />

13 13.375 12.502 12.285 12.025 11.707 11.545<br />

14 14.000 13.086 12.859 12.586 12.253 12.086<br />

16 16.000 14.967 14.696 14.385 14.005 13.812<br />

18 18.000 16.826 16.533 16.183 15.755 15.539<br />

20 20.000 18.696 18.370 17.982 17.507 17.265<br />

22 22.000 20.565 20.206 19.778 19.257 18.992<br />

24 24.000 22.435 22.043 21.577 21.007 20.718<br />

26 26.000 24.304 23.880 23.375 22.759 22.445<br />

28 28.000 26.173 25.717 25.174 24.508 24.171<br />

30 30.000 28.043 27.554 26.971 26.258 25.898<br />

32M 31.594 29.541 29.054 28.414 27.663 27.288<br />

36 36.000 33.651 33.054 32.366 31.510 31.075<br />

40M 39.469 35.898 36.225 35.469 34.561<br />

42 42.000 39.261 38.576 37.760 36.761<br />

48M 47.382 44.302 43.526 42.616 41.489<br />

U = 0.8492 CR H 0.63 S 0.54 in m/s (2-22)<br />

All parameters in Equation 2-22 must be in SI units. There is no consistency in using the<br />

Hazen–Williams formula from small to large pi<strong>pe</strong>s.<br />

Despite the fact that commercial publications from pi<strong>pe</strong> suppliers sometimes use<br />

Hazen–Williams equations to determine pressure loss of slurries, this method is highly<br />

discouraged for long pi<strong>pe</strong>lines.<br />

2.4 THE HYDRAULIC FRICTION GRADIENT<br />

OF WATER IN RUBBER-LINED STEEL PIPES<br />

Equations 2-15 to 2-19 have established a relationship between the friction factor and the<br />

Reynolds number. To compute the latter, the pro<strong>pe</strong>rties of water as a carrier fluid are presented<br />

in Tables 2-7 and 2-8.


2.20 CHAPTER TWO<br />

TABLE 2-5 Dimensions of North American HDPE Pi<strong>pe</strong>s for Pressure Ratings 128 psi<br />

to 300 psi at a Tem<strong>pe</strong>rature of 23°C (73.4 °F)<br />

DR 13.5 DR11 DR 9 DR 7.3 DR6.3<br />

Pressure rating 128 psi 160 psi 200 psi 254 psi 300 psi<br />

Average Average Average Average Average Average<br />

outside inside inside inside inside inside<br />

Pi<strong>pe</strong> size diameter diameter diameter diameter diameter diameter<br />

(inch) (inch) (inch) (inch) (inch) (inch) (inch)<br />

3 3.500 2.951 2.826 2.675 2.485 2.321<br />

4 4.500 3.795 3.633 3.440 3.194 2.986<br />

5 5.563 4.690 4.490 4.253 3.948 3.690<br />

6 6.625 5.584 5.349 5.065 4.700 4.395<br />

7 7.125 6.006 5.751 5.446 5.056 4.727<br />

8 8.625 7.270 6.693 6.594 6.119 5.723<br />

10 10.750 9.062 8.679 8.219 7.627 7.133<br />

12 12.750 10.749 10.293 9.745 9.046 8.459<br />

13 13.375 11.274 10.797 10.225 9.491 8.674<br />

14 14.000 11.802 11.301 10.701 9.934 9.289<br />

16 16.000 13.488 12.915 12.231 11.853 10.615<br />

18 18.000 15.174 14.532 13.760 12.772 —<br />

20 20.000 16.880 16.145 15.289 14.191 —<br />

22 22.000 18.544 17.780 16.819 — —<br />

24 24.000 20.231 19.374 18.346<br />

26 26.000 21.917 20.988 19.875<br />

28 28.000 23.603 22.606 21.405<br />

30 30.000 25.289 24.219 22.934<br />

32M 31.594 26.645 25.527<br />

TABLE 2-6 Hazen–Williams Roughness Coefficients<br />

Description Roughness coefficient<br />

Extremely smooth pi<strong>pe</strong> 140<br />

Very smooth pi<strong>pe</strong> 130<br />

Concrete pi<strong>pe</strong> 120<br />

Riveted new pi<strong>pe</strong>s and tiled channels 110<br />

Normal cast pi<strong>pe</strong>s, 10 year old steel pi<strong>pe</strong>s, masonry channels 100<br />

Very rough pi<strong>pe</strong>s 60


FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

TABLE 2-7 Physical Pro<strong>pe</strong>rties of Water in SI Units<br />

2.21<br />

Dynamic Kinematic Surface Vapor<br />

viscosity, viscosity, tension pressure<br />

kinematic dynamic in contact at atmospheric<br />

Tem<strong>pe</strong>rature Density �L viscosity � viscosity � with air � pressure<br />

T (°C) (kg/m3 ) (mPa·s) (km2 /s) (N/m) (kN/m2 )<br />

0 999.8 1.781 1.785 0.0756 0.61<br />

5 1000 1.518 1.519 0.0749 0.87<br />

10 999.7 1.307 1.306 0.0742 1.23<br />

15 999.1 1.139 1.139 0.0735 1.70<br />

20 998.2 1.002 1.003 0.0728 2.34<br />

25 997.0 0.890 0.893 0.0720 3.17<br />

30 995.7 0.798 0.800 0.0712 4.24<br />

40 992.2 0.653 0.658 0.0696 7.38<br />

50 988.0 0.547 0.553 0.0679 12.33<br />

60 983.2 0.466 0.474 0.0662 19.92<br />

70 977.8 0.404 0.413 0.0644 31.16<br />

80 971.8 0.354 0.364 0.0626 47.34<br />

90 965.3 0.315 0.326 0.0608 70.10<br />

100 958.4 0.282 0.294 0.0589 101.33<br />

TABLE 2-8 Physical Pro<strong>pe</strong>rties of Water in USCS Units<br />

Vapor pressure<br />

Tem<strong>pe</strong>rature Density Kinematic Dynamic Surface tension in at atmospheric<br />

T � L viscosity � viscosity � contact with air � pressure<br />

(°F) (slug/ft 3 ) (lbf-sec/ft 2 ) (ft 2 /sec) (lbf/ft) (psia)<br />

32 1.940 0.00003746 0.00001931 0.00518 0.09<br />

40 1.940 0.00003229 0.00001664 0.00614 0.12<br />

50 1.940 0.00002735 0.00001410 0.00509 0.18<br />

60 1.938 0.00002359 0.00001217 0.00504 0.26<br />

70 1.936 0.0000205 0.00001059 0.00498 0.36<br />

80 1.934 0.00001799 0.00009300 0.00492 0.51<br />

90 1.931 0.00001595 0.00008260 0.00486 0.70<br />

100 1.927 0.00001424 0.00007390 0.00480 0.95<br />

110 1.923 0.00001284 0.00006670 0.00473 1.27<br />

120 1.918 0.00001168 0.00006090 0.00467 1.69<br />

130 1.913 0.00001069 0.00005580 0.00460 2.22<br />

140 1.908 0.00000981 0.00005140 0.00454 2.89<br />

150 1.902 0.00000905 0.00004760 0.00447 3.72<br />

160 1.896 0.00000838 0.00004420 0.00441 4.74<br />

170 1.890 0.0000078 0.00004130 0.00434 5.99<br />

180 1.883 0.00000726 0.00003850 0.00427 7.51<br />

190 1.876 0.00000678 0.00003620 0.00420 9.34<br />

200 1.868 0.00000637 0.00003410 0.00413 11.52<br />

212 1.860 0.00000593 0.00003190 0.00404 14.70


2.22 CHAPTER TWO<br />

The friction losses are expressed by the following equation:<br />

H = (2-23)<br />

where<br />

H = head due to losses (in meters for SI units, in ft for USCS units)<br />

L = length of the pi<strong>pe</strong> (in meters for SI units, in ft for USCS units)<br />

fD = Darcy friction factor<br />

DI = pi<strong>pe</strong> inner diameter<br />

V = s<strong>pe</strong>ed of flow in the pi<strong>pe</strong> (m/s or ft/s)<br />

g = acceleration due to gravity (9.81 m/s or 32.2 ft/s)<br />

The hydraulic friction gradient is defined as the head loss <strong>pe</strong>r unit length. It is defined as<br />

fDV iw = = (2-24)<br />

This is a very important parameter that will be used in Chapter 4 to evaluate the friction<br />

loss of Newtonian flows.<br />

Calculations of the friction factor by Equation 2-18 require iterations. The Moody diagram<br />

is a logarithmic scale that is rather difficult to use and prone to reading errors. With<br />

modern computers, a simple program will give more accurate numbers for the Darcy friction<br />

factor than from reading the Moody curve on a difficult logarithmic scale. The following<br />

program was written for plain steel and rubber-lined steel pi<strong>pe</strong>s. It uses standard<br />

US pi<strong>pe</strong> sizes and applies the Swamee–Jain equation (2-19).<br />

2<br />

fDV H<br />

� �<br />

L 2gDI<br />

2L �<br />

2gDI<br />

DIM PIP(300), t(300), a(300), q(300), ep(300)<br />

pi = 4 * ATN(1)<br />

DEF fnlog10 (X) = LOG(X) * .4342944<br />

‘ p refers to nominal size<br />

‘ outside diameter is stated<br />

p2 = 2.375<br />

p3 = 3.5<br />

p4 = 4.5<br />

p5 = 5.633<br />

p6 = 6.625<br />

p8 = 8.625<br />

p10 = 10.75<br />

p12 = 12.75<br />

p14 = 14<br />

p16 = 16<br />

p18 = 18<br />

p20 = 20<br />

p24 = 24<br />

p30 = 30<br />

p36 = 36<br />

p42 = 42<br />

p48 = 48<br />

INPUT “choose between steel (1) and rubber (2)”, ch<br />

IF ch = 1 THEN tr1 = 0<br />

IF ch = 1 THEN tr2 = 0


IF ch = 2 THEN tr1 = .254<br />

IF ch = 2 THEN tr2 = .375<br />

IF ch = 1 THEN e = .00006<br />

IF ch = 2 THEN e = .00015<br />

ef = e / .0254<br />

IF ch = 1 THEN LPRINT “steel pi<strong>pe</strong>”<br />

IF ch = 2 THEN LPRINT “rubber lined pi<strong>pe</strong>”<br />

‘2<br />

FOR I = 1 TO 4<br />

t(1) = .154<br />

t(2) = .218<br />

t(3) = .344<br />

t(4) = .436<br />

PIP(I) = p2 - 2 * t(I) - 2 * tr1<br />

ep(I) = ef / PIP(I)<br />

LPRINT USING “p<strong>pe</strong> od = ##.### in pip id = ##.#### in thick =<br />

#.### in k/d =<br />

#.########”; p2; PIP(I); t(I); ep(I)<br />

GOSUB swamee<br />

NEXT I<br />

LPRINT<br />

LPRINT<br />

The program is re<strong>pe</strong>ated for all US sizes of pi<strong>pe</strong>s (not shown here)<br />

swamee:<br />

d1 = PIP(I) * .0254<br />

a = .25 * pi * d1 ^ 2<br />

FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

FOR K = 1 TO 5<br />

q = a * K * 1000<br />

qus = (q * 60 / 3.7854)<br />

fd = .4<br />

em = ep(I)<br />

Re = 1000 * K * d1 / .001<br />

110<br />

z = (em / 3.7) + (5.74 / Re ^ .9)<br />

y = fnlog10(z)<br />

fd = .25 / y ^ 2<br />

1111 hl = fd * K ^ 2 / (2 * 9.81 * d1)<br />

PRINT “revised swamee factor “; fd<br />

LPRINT USING “veloc = ##.### m/s; flow q = ####.#### L/s;<br />

flow = ######.## gpm “; K;<br />

q; qus<br />

LPRINT USING “RE = #########; fd = #.#####; hL =<br />

#####.####”; Re; fd; hl<br />

LPRINT<br />

PRINT “iteration error in swamee friction factor “; dg<br />

2.23


2.24 CHAPTER TWO<br />

TABLE 2-9 Flow and Hydraulic Friction Gradient for Steel Pi<strong>pe</strong>s at a S<strong>pe</strong>ed of 1 m/s to<br />

5m/s (3.28 to 16.4 ft/sec) for Water at a Density of 1000 kg/m 3 (62.3 lbs/ft 3 ) and a<br />

Kinematic Viscosity of 1 cP (2.09 × 10 –5 lbf-sec/ft 2 ) Using the Swamee–Jain Equation*<br />

Relative Friction<br />

Wall roughness S<strong>pe</strong>ed S<strong>pe</strong>ed Flow Darcy gradient<br />

Denomination thickness for new of flow Flow of flow US Reynolds friction (m/m)<br />

(inch) (inch) pi<strong>pe</strong> (m/s) L/s (ft/s) gpm number factor or (ft/ft)<br />

2� Sch 40<br />

2� Sch 80<br />

2� Sch 160<br />

3� Sch 40<br />

3� Sch 80<br />

3� Sch 160<br />

4� Sch 40<br />

4� Sch 80<br />

4� Sch 160<br />

0.154<br />

0.218<br />

0.344<br />

0.216<br />

0.300<br />

0.438<br />

0.237<br />

0.337<br />

0.531<br />

0.001143<br />

0.00121<br />

0.001400<br />

0.000770<br />

0.000815<br />

0.000900<br />

0.000587<br />

0.000617<br />

0.000687<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

2.2<br />

4.3<br />

6.5<br />

8.6<br />

10.8<br />

1.9<br />

3.8<br />

5.7<br />

7.6<br />

9.5<br />

1.4<br />

2.9<br />

4.3<br />

5.8<br />

7.2<br />

4.77<br />

9.5<br />

14.3<br />

19.08<br />

23.8<br />

4.3<br />

8.5<br />

12.8<br />

17.1<br />

21.3<br />

3.5<br />

7<br />

10.5<br />

13.9<br />

17.5<br />

8.21<br />

16.4<br />

24.6<br />

32.8<br />

41.1<br />

7.4<br />

14.8<br />

22.3<br />

29.7<br />

37.1<br />

6<br />

12<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

34.3<br />

69<br />

103<br />

137<br />

172<br />

30.2<br />

60.4<br />

90.6<br />

121<br />

151<br />

22.9<br />

45.7<br />

68.6<br />

91.4<br />

114<br />

75.6<br />

151<br />

227<br />

303<br />

378<br />

67.5<br />

135<br />

203<br />

270<br />

337<br />

55.3<br />

111<br />

166<br />

221<br />

277<br />

130.2<br />

260.4<br />

391<br />

521<br />

651<br />

117<br />

253<br />

353<br />

470<br />

588<br />

95<br />

190<br />

52,502<br />

105,004<br />

157505<br />

210007<br />

262509<br />

49,251<br />

98,501<br />

147,752<br />

197002<br />

246253<br />

42,850<br />

85,700<br />

128,549<br />

171,399<br />

214249<br />

77,927<br />

155,854<br />

233,782<br />

311709<br />

389636<br />

73,660<br />

147,320<br />

202,980<br />

294,640<br />

368,300<br />

66,650<br />

133,299<br />

199,949<br />

266,598<br />

332,248<br />

102,260<br />

204,521<br />

306,781<br />

409,042<br />

511,302<br />

97,180<br />

194,361<br />

291,541<br />

388,722<br />

485,902<br />

87,325<br />

174,650<br />

0.0244<br />

0.0227<br />

0.0221<br />

0.0217<br />

0.0214<br />

0.0248<br />

0.0231<br />

0.0224<br />

0.0220<br />

0.0218<br />

0.0258<br />

0.0239<br />

0.0232<br />

0.0228<br />

0.023<br />

0.02213<br />

0.0206<br />

0.0200<br />

0.0197<br />

0.0195<br />

0.0224<br />

0.0209<br />

0.0203<br />

0.0199<br />

0.0197<br />

0.0230<br />

0.0214<br />

0.0208<br />

0.0204<br />

0.0202<br />

0.0207<br />

0.0193<br />

0.0188<br />

0.0185<br />

0.0183<br />

0.021<br />

0.01957<br />

0.019<br />

0.0187<br />

0.0185<br />

0.0215<br />

0.0201<br />

0.0237<br />

0.0883<br />

0.1927<br />

0.3367<br />

0.5204<br />

0.0257<br />

0.0956<br />

0.2087<br />

0.3648<br />

0.5637<br />

0.0306<br />

0.114<br />

0.249<br />

0.434<br />

0.671<br />

0.0145<br />

0.054<br />

0.118<br />

0.206<br />

0.319<br />

0.0155<br />

0.0579<br />

0.1264<br />

0.2209<br />

0.3415<br />

0.0176<br />

0.0655<br />

0.1431<br />

0.2501<br />

0.3866<br />

0.0103<br />

0.0386<br />

0.0843<br />

0.1473<br />

0.2278<br />

0.011<br />

0.041<br />

0.09<br />

0.157<br />

0.243<br />

0.0126<br />

0.047


TABLE 2-9 Continued<br />

FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

2.25<br />

Relative Friction<br />

Wall roughness S<strong>pe</strong>ed S<strong>pe</strong>ed Flow Darcy gradient<br />

Denomination thickness for new of flow Flow of flow US Reynolds friction (m/m)<br />

(inch) (inch) pi<strong>pe</strong> (m/s) L/s (ft/s) gpm number factor or (ft/ft)<br />

4� Sch 160<br />

5� Sch 40<br />

5� Sch 80<br />

5� Sch 160<br />

6� Sch 40<br />

6� Sch 80<br />

6� Sch 160<br />

8� Sch 40<br />

8� Sch 80<br />

0.257<br />

0.375<br />

0.625<br />

0.280<br />

0.432<br />

0.719<br />

0.322<br />

0.500<br />

0.000461<br />

0.000484<br />

0.000539<br />

0.000389<br />

0.000410<br />

0.000455<br />

0.000296<br />

0.000310<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

18<br />

24<br />

30<br />

13.3<br />

26.5<br />

39.8<br />

53<br />

66.3<br />

12.1<br />

24.1<br />

36.2<br />

48.4<br />

60.4<br />

9.7<br />

19.5<br />

29.2<br />

38.9<br />

48.7<br />

18.7<br />

37.3<br />

55.9<br />

74.6<br />

93.2<br />

16.8<br />

33.6<br />

50.5<br />

67.3<br />

84<br />

13.6<br />

27.3<br />

40.9<br />

54.5<br />

68.2<br />

32.3<br />

65.6<br />

96.8<br />

129.1<br />

161.4<br />

29.4<br />

58.9<br />

88.4<br />

117.8<br />

147.3<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

285<br />

380<br />

475<br />

210<br />

421<br />

631<br />

841<br />

1052<br />

192<br />

383<br />

575<br />

766<br />

958<br />

154.3<br />

308.6<br />

463<br />

617<br />

772<br />

295.4<br />

591<br />

886<br />

1182<br />

1477<br />

267<br />

533<br />

800<br />

1066<br />

1333<br />

216<br />

432<br />

648<br />

864<br />

1,080<br />

512<br />

1,023<br />

1,535<br />

2,046<br />

2,558<br />

467<br />

934<br />

1,401<br />

1,868<br />

2335<br />

261,976<br />

349,301<br />

436,626<br />

129,972<br />

259,944<br />

389,915<br />

519,887<br />

649,859<br />

124,028<br />

248,056<br />

372,085<br />

496,113<br />

620,141<br />

111,328<br />

222,656<br />

333,985<br />

445,313<br />

556,641<br />

154,051<br />

308,102<br />

462,153<br />

616,204<br />

770,255<br />

146,329<br />

292,659<br />

438,988<br />

585,318<br />

731,647<br />

131,750<br />

263,500<br />

395,249<br />

526,999<br />

658,749<br />

202,717<br />

405,435<br />

608,152<br />

810,870<br />

1,013,587<br />

193,675<br />

387,350<br />

581,025<br />

774,700<br />

968,375<br />

0.0195<br />

0.0192<br />

0.019<br />

0.0196<br />

0.0183<br />

0.0178<br />

0.0175<br />

0.0173<br />

0.1981<br />

0.0185<br />

0.018<br />

0.0177<br />

0.0175<br />

0.0203<br />

0.0189<br />

0.0184<br />

0.0181<br />

0.0179<br />

0.0189<br />

0.0176<br />

0.0171<br />

0.0169<br />

0.0167<br />

0.0191<br />

0.0178<br />

0.0173<br />

0.0170<br />

0.0169<br />

0.0195<br />

0.0183<br />

0.0177<br />

0.0174<br />

0.0173<br />

0.0177<br />

0.0166<br />

0.0161<br />

0.0159<br />

0.0157<br />

0.0179<br />

0.0168<br />

0.0163<br />

0.0160<br />

0.0158<br />

0.102<br />

0.179<br />

0.277<br />

0.0077<br />

0.0287<br />

0.0628<br />

0.1098<br />

0.1697<br />

0.0081<br />

0.0304<br />

0.0665<br />

0.1163<br />

0.1797<br />

0.0093<br />

0.0347<br />

0.0759<br />

0.1327<br />

0.2052<br />

0.0062<br />

0.233<br />

0.051<br />

0.0892<br />

0.138<br />

0.0066<br />

0.0248<br />

0.0543<br />

0.095<br />

0.147<br />

0.0076<br />

0.0282<br />

0.0617<br />

0.108<br />

0.167<br />

0.0045<br />

0.0167<br />

0.0365<br />

0.0639<br />

0.0988<br />

0.0047<br />

0.0176<br />

0.0386<br />

0.0675<br />

0.1044<br />

(continued)


2.26 CHAPTER TWO<br />

TABLE 2-9 Continued<br />

Relative Friction<br />

Wall roughness S<strong>pe</strong>ed S<strong>pe</strong>ed Flow Darcy gradient<br />

Denomination thickness for new of flow Flow of flow US Reynolds friction (m/m)<br />

(inch) (inch) pi<strong>pe</strong> (m/s) L/s (ft/s) gpm number factor or (ft/ft)<br />

8� Sch 160<br />

10� Sch 40 S<br />

10� Sch 60 X<br />

10� Sch 120 XX<br />

12� S<br />

12� X<br />

12� Sch 120 XX<br />

14� S<br />

14� X<br />

0.906<br />

0.365<br />

0.500<br />

1.000<br />

0.375<br />

0.500<br />

1.000<br />

0.375<br />

0.500<br />

0.00347<br />

0.000236<br />

0.000243<br />

0.000269<br />

.000197<br />

0.000201<br />

0.00022<br />

0.000178<br />

0.000182<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

23.5<br />

47<br />

70.6<br />

94.1<br />

117.6<br />

51<br />

102<br />

153<br />

204<br />

254<br />

48<br />

96<br />

145<br />

193<br />

241<br />

39<br />

78<br />

116<br />

155<br />

194<br />

73<br />

146<br />

219<br />

292<br />

365<br />

59<br />

117<br />

177<br />

235<br />

293<br />

59<br />

117<br />

177<br />

235<br />

293<br />

89<br />

178<br />

267<br />

356<br />

445<br />

86<br />

171<br />

257<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

373<br />

746<br />

1,118<br />

1,491<br />

1,864<br />

806<br />

1,613<br />

2,419<br />

3,226<br />

4,032<br />

764<br />

1,527<br />

2,291<br />

3,054<br />

3,818<br />

615<br />

1,230<br />

1,845<br />

2,460<br />

3,075<br />

1,157<br />

2,313<br />

3,470<br />

4,626<br />

5,783<br />

928<br />

1,856<br />

2,784<br />

3,713<br />

4,641<br />

928<br />

1856<br />

2784<br />

3713<br />

4641<br />

1,410<br />

2,820<br />

4,231<br />

5,640<br />

7,050<br />

1,357<br />

2,175<br />

4,072<br />

173,050<br />

346,100<br />

519,151<br />

692,201<br />

865,251<br />

254,508<br />

509,106<br />

763,524<br />

1,018,032<br />

1,272,540<br />

247,650<br />

495,300<br />

742,950<br />

990,600<br />

1,238,250<br />

222,250<br />

444,500<br />

666,750<br />

889,000<br />

1,111,250<br />

304,800<br />

609,600<br />

914,400<br />

1,219,200<br />

1,524,000<br />

273,050<br />

546,100<br />

819,150<br />

1,099,000<br />

1,365,250<br />

273,050<br />

546,100<br />

819,150<br />

1,099,000<br />

1,365,250<br />

336,550<br />

673,100<br />

1,009,650<br />

1,346,200<br />

1,682,750<br />

330,200<br />

660,400<br />

990,600<br />

0.0184<br />

0.0172<br />

0.0167<br />

0.0164<br />

0.0163<br />

0.0169<br />

0.0158<br />

0.0154<br />

0.0151<br />

0.0150<br />

0.017<br />

0.0159<br />

0.0155<br />

0.0152<br />

0.0151<br />

0.0174<br />

0.0163<br />

0.0158<br />

0.0156<br />

0.0154<br />

0.0163<br />

0.0152<br />

0.0148<br />

0.0146<br />

0.0144<br />

0.0166<br />

0.0156<br />

0.0152<br />

0.0149<br />

0.0148<br />

0.0166<br />

0.0156<br />

0.0152<br />

0.0149<br />

0.0148<br />

0.0159<br />

0.0149<br />

0.0145<br />

0.0143<br />

0.0141<br />

0.016<br />

0.015<br />

0.0146<br />

0.0054<br />

0.0202<br />

0.0443<br />

0.0774<br />

0.1197<br />

0.0034<br />

0.0127<br />

0.0277<br />

0.0485<br />

0.0750<br />

0.0035<br />

0.0131<br />

0.0286<br />

0.0501<br />

0.0775<br />

0.004<br />

0.0149<br />

0.0327<br />

0.0571<br />

0.0883<br />

0.0027<br />

0.0102<br />

0.0223<br />

0.0390<br />

0.0603<br />

0.0031<br />

0.0116<br />

0.0255<br />

0.0445<br />

0.0689<br />

0.0031<br />

0.0116<br />

0.0255<br />

0.0445<br />

0.0689<br />

0.0024<br />

0.0090<br />

0.0198<br />

0.0346<br />

0.0536<br />

0.0025<br />

0.0093<br />

0.0202


TABLE 2-9 Continued<br />

FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

2.27<br />

Relative Friction<br />

Wall roughness S<strong>pe</strong>ed S<strong>pe</strong>ed Flow Darcy gradient<br />

Denomination thickness for new of flow Flow of flow US Reynolds friction (m/m)<br />

(inch) (inch) pi<strong>pe</strong> (m/s) L/s (ft/s) gpm number factor or (ft/ft)<br />

14� X<br />

14� Sch 120<br />

16� Sch 30 S<br />

16� X<br />

16� Sch 120<br />

18� S<br />

18� X<br />

18� Sch 120<br />

20� Sch 20S<br />

1.062<br />

0.375<br />

0.500<br />

1.291<br />

0.375<br />

0.500<br />

1.375<br />

0.375<br />

0.000531<br />

0.000155<br />

0.0001575<br />

0.000176<br />

0.000137<br />

0.000139<br />

0.000155<br />

0.000123<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

343<br />

428<br />

72<br />

143<br />

214<br />

286<br />

357<br />

118<br />

236<br />

354<br />

471<br />

589<br />

114<br />

228<br />

342<br />

456<br />

570<br />

91<br />

182<br />

274<br />

365<br />

456<br />

151<br />

302<br />

452<br />

603<br />

754<br />

146<br />

293<br />

439<br />

586<br />

732<br />

118<br />

236<br />

354<br />

471<br />

589<br />

188<br />

375<br />

563<br />

751<br />

939<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

5,429<br />

6,787<br />

1,133<br />

2,266<br />

3398<br />

4531<br />

5664<br />

1,868<br />

3,736<br />

5,604<br />

7,471<br />

9,339<br />

1,807<br />

3,614<br />

5,421<br />

7,228<br />

9,035<br />

1,446<br />

2,892<br />

4,338<br />

5784<br />

7230<br />

2,390<br />

4,780<br />

7,170<br />

9,560<br />

11,949<br />

2,321<br />

4,642<br />

6,963<br />

9,284<br />

11,605<br />

1,868<br />

3,734<br />

5,604<br />

7,471<br />

9340<br />

2,976<br />

5,952<br />

8,929<br />

11,905<br />

14,881<br />

1,320,800<br />

1,651,000<br />

301,650<br />

603,301<br />

904,951<br />

1,206,602<br />

1,508,251<br />

387,350<br />

774,700<br />

1,162,050<br />

1,549,400<br />

1,936,750<br />

381,000<br />

762,000<br />

1,143,000<br />

1,524,000<br />

1,905,000<br />

340,817<br />

681,634<br />

1,022,452<br />

1,363,269<br />

1,704,086<br />

438,150<br />

876,300<br />

1,314,450<br />

1,752,600<br />

2,190,750<br />

431,800<br />

863,600<br />

1,295,400<br />

1,727,200<br />

2,159,000<br />

387,350<br />

774,700<br />

1,162,050<br />

1,549,000<br />

1,966,750<br />

488,950<br />

977,900<br />

1,466,850<br />

1,955,800<br />

2,444,750<br />

0.0144<br />

0.0142<br />

0.0163<br />

0.0153<br />

0.01485<br />

0.0146<br />

0.0145<br />

0.0155<br />

0.0145<br />

0.0141<br />

0.0139<br />

0.0138<br />

0.0155<br />

0.0146<br />

0.0142<br />

0.0139<br />

0.0138<br />

0.0159<br />

0.0149<br />

0.0145<br />

0.0143<br />

0.0141<br />

0.0151<br />

0.0142<br />

0.0138<br />

0.0136<br />

0.0134<br />

0.0151<br />

0.0142<br />

0.0138<br />

0.0136<br />

0.0135<br />

0.0155<br />

0.0145<br />

0.0141<br />

0.0139<br />

0.0138<br />

0.0148<br />

0.0139<br />

0.0135<br />

0.0133<br />

0.0131<br />

0.0354<br />

0.0548<br />

0.0028<br />

0.0103<br />

0.0226<br />

0.0395<br />

0.0611<br />

0.002<br />

0.0076<br />

0.0167<br />

0.0292<br />

0.0452<br />

0.0021<br />

0.0078<br />

0.0170<br />

0.0298<br />

0.0461<br />

0.0024<br />

0.0089<br />

0.0195<br />

0.0341<br />

0.0527<br />

0.0018<br />

0.0066<br />

0.0144<br />

0.0252<br />

0.0390<br />

0.0018<br />

0.0067<br />

0.0147<br />

0.0257<br />

0.0397<br />

0.0020<br />

0.0076<br />

0.0167<br />

0.0292<br />

0.0452<br />

0.0015<br />

0.0058<br />

0.0126<br />

0.0221<br />

0.0342<br />

(continued)


2.28 CHAPTER TWO<br />

TABLE 2-9 Continued<br />

Relative Friction<br />

Wall roughness S<strong>pe</strong>ed S<strong>pe</strong>ed Flow Darcy gradient<br />

Denomination thickness for new of flow Flow of flow US Reynolds friction (m/m)<br />

(inch) (inch) pi<strong>pe</strong> (m/s) L/s (ft/s) gpm number factor or (ft/ft)<br />

20� Sch 30 X<br />

24� Sch 20 S<br />

24� Sch S<br />

0.500<br />

0.375<br />

0.500<br />

0.000124<br />

0.000102<br />

0.000102<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

183<br />

366<br />

549<br />

732<br />

915<br />

274<br />

548<br />

822<br />

1096<br />

1370<br />

268<br />

536<br />

804<br />

1072<br />

1340<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

*Dimensions are based on ANSI B36.1 and B31.1. Absolute roughness of steel is taken as 60 �m.<br />

‘INPUT “hit any key to continue”; m$<br />

CLS<br />

NEXT K<br />

RETURN<br />

2,899<br />

5,799<br />

8,698<br />

11,598<br />

14,497<br />

4,341<br />

8,683<br />

13,025<br />

17,366<br />

21,707<br />

4,249<br />

8,497<br />

12,746<br />

16,995<br />

21,243<br />

482,600<br />

965,200<br />

1,447,800<br />

1,930,400<br />

2,413,000<br />

590,550<br />

1,181,100<br />

1,771,650<br />

2,362,200<br />

2,952,750<br />

584,200<br />

1,1684,000<br />

1,752,600<br />

2,336,800<br />

2,921,000<br />

0.0148<br />

0.0139<br />

0.0135<br />

0.0133<br />

0.0132<br />

0.0142<br />

0.0134<br />

0.013<br />

0.0128<br />

0.01267<br />

0.01425<br />

0.01338<br />

0.01302<br />

0.01282<br />

0.01269<br />

0.0016<br />

0.0059<br />

0.0128<br />

0.0225<br />

0.0348<br />

0.0012<br />

0.0046<br />

0.0101<br />

0.0177<br />

0.0273<br />

0.0012<br />

0.0047<br />

0.0102<br />

0.0179<br />

0.0277<br />

The range of s<strong>pe</strong>eds used to carry solids in Newtonian flows is typically between 1.5<br />

m/s and 5m/s.<br />

Tables 2-9 and 2-10 present friction factor and head losses for water as a carrier fluid<br />

for plain and rubber-lined steel pi<strong>pe</strong>s. There is no point in tabulating other fluids here as<br />

they are rarely used for <strong>slurry</strong> mixtures.<br />

The hydraulic friction gradient of water in rubber-lined pi<strong>pe</strong>s in the range of 2� to 18�<br />

is presented in Figures 2-5 to 2-13. Rubber thickness of 6.4 mm (0.25�) was assumed for<br />

2�, 3�, 4�, and 6� (up to 150 mm) pi<strong>pe</strong>s. Rubber thickness of 9.5 mm (0.375�) was assumed<br />

for 8� to 24� (200 to 610 mm NB) pi<strong>pe</strong>s. HDPE friction head was plotted for similar sizes<br />

at SDR11 (suitable for 100 psi pressure), to mark the advantages of reduced friction at<br />

these sizes using HDPE instead of rubber-lined pi<strong>pe</strong>s, wherever it may be appropriate.<br />

The design engineer must take in account the pressure limitations of HDPE pi<strong>pe</strong>s versus<br />

rubber-lined steel pi<strong>pe</strong>s.<br />

The hydraulic friction gradient for HDPE pi<strong>pe</strong>s up to a size of 20� (508 mm), and for<br />

s<strong>pe</strong>eds in the range of 1 to 5 m/s (3.3 to 16.5 ft/sec) is presented in Table 2-11<br />

These curves and tables allow an easier and accurate determination of the hydraulic<br />

friction gradient of water than the Moody diagram.


FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

2.29<br />

TABLE 2-10 Flow and Hydraulic Friction Gradients for Rubber-Lined Steel Pi<strong>pe</strong>s at a<br />

S<strong>pe</strong>ed of 1 m/s to 5 m/s (3.28 to 16.4 ft/sec) for Water at a Density of 1000 kg/m 3 (62.3<br />

lbs/ft 3 ) and a Kinematic Viscosity of 1 cP (2.09 × 10 –5 lbf-sec/ft 2 ) Using the Swamee–Jain<br />

Equation*<br />

Relative Energy<br />

Wall roughness S<strong>pe</strong>ed S<strong>pe</strong>ed Flow Darcy gradient<br />

Pi<strong>pe</strong> size thickness for new of flow Flow of flow US Reynolds friction (m/m)<br />

(inch) (inch) pi<strong>pe</strong> (m/s) L/s (ft/s) gpm number factor or (ft/ft)<br />

2� Sch 40<br />

2� Sch 80<br />

2� Sch<br />

160<br />

3� Sch 40<br />

3� Sch 80<br />

3� Sch 160<br />

4� Sch 40<br />

4� Sch 80<br />

Steel 0.154<br />

Rubber 0.250<br />

Steel 0.218<br />

Rubber 0.250<br />

Steel 0.344<br />

Rubber 0.250<br />

Steel 0.216<br />

Rubber 0.250<br />

Steel 0.300<br />

Rubber 0.250<br />

Steel 0.438<br />

Rubber 0.250<br />

Steel 0.237<br />

Rubber 0.250<br />

Steel 0.337<br />

Rubber 0.250<br />

0.003788<br />

0.004123<br />

0.005008<br />

0.002487<br />

0.002791<br />

0.001679<br />

0.00178<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1.2<br />

2.5<br />

3.7<br />

4.9<br />

6.2<br />

1<br />

2.1<br />

3.1<br />

4.2<br />

5.2<br />

0.7<br />

1.4<br />

2.1<br />

2.8<br />

3.5<br />

3.3<br />

6.6<br />

10<br />

13.3<br />

16.6<br />

2.9<br />

5.8<br />

8.7<br />

11.6<br />

14.5<br />

2.3<br />

4.5<br />

6.8<br />

9.1<br />

11.3<br />

6.3<br />

12.5<br />

18.8<br />

25.1<br />

31.4<br />

5.6<br />

11.2<br />

16.7<br />

23.3<br />

27.8<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

19.5<br />

39<br />

58.6<br />

78<br />

97.6<br />

16.5<br />

33<br />

49.4<br />

65.8<br />

82.2<br />

11.2<br />

22.3<br />

33.5<br />

44.7<br />

55.8<br />

53<br />

105<br />

158<br />

211<br />

263<br />

46<br />

92<br />

138<br />

184<br />

230<br />

36<br />

72<br />

108<br />

144<br />

180<br />

99<br />

199<br />

298<br />

398<br />

497<br />

88.4<br />

177<br />

265<br />

354<br />

442<br />

39,599<br />

79,197<br />

118,796<br />

158,394<br />

197,993<br />

36,347<br />

72,695<br />

109,042<br />

145,390<br />

181,737<br />

29,947<br />

59,893<br />

89,840<br />

119,786<br />

149,733<br />

65,024<br />

130,048<br />

195,072<br />

260,096<br />

325,120<br />

60,757<br />

121,514<br />

182,270<br />

243,027<br />

303,784<br />

53,746<br />

107,493<br />

161,239<br />

214,986<br />

268,732<br />

89,357<br />

178,714<br />

268,072<br />

357,429<br />

446,786<br />

84,277<br />

168,554<br />

252,832<br />

337,109<br />

421,386<br />

0.0309<br />

0.0297<br />

0.289<br />

0.0287<br />

0.0287<br />

0.0317<br />

0.0299<br />

0.0296<br />

0.0295<br />

0.0337<br />

0.0337<br />

0.0322<br />

0.0317<br />

0.0314<br />

0.0312<br />

0.0269<br />

0.0257<br />

0.0253<br />

0.0251<br />

0.0250<br />

0.0274<br />

0.0262<br />

0.0258<br />

0.0256<br />

0.0255<br />

0.0283<br />

0.0272<br />

0.0267<br />

0.0265<br />

0.0263<br />

0.0247<br />

0.0237<br />

0.0233<br />

0.0231<br />

0.0229<br />

0.0251<br />

0.0240<br />

0.0236<br />

0.0234<br />

0.0233<br />

0.0399<br />

0.153<br />

0.338<br />

0.595<br />

0.925<br />

0.045<br />

0.171<br />

0.377<br />

0.665<br />

1.033<br />

0.0574<br />

0.2195<br />

0.4854<br />

0.8551<br />

1.3284<br />

0.0211<br />

0.0808<br />

0.1788<br />

0.3150<br />

0.4895<br />

0.0230<br />

0.088<br />

0.1949<br />

0.3435<br />

0.5337<br />

0.0269<br />

0.103<br />

0.228<br />

0.402<br />

0.624<br />

0.0141<br />

0.054<br />

0.1195<br />

0.2106<br />

0.3273<br />

0.0151<br />

0.0581<br />

0.1287<br />

0.2268<br />

0.352<br />

(continued)


2.30 CHAPTER TWO<br />

TABLE 2-10 Continued<br />

Relative Energy<br />

Wall roughness S<strong>pe</strong>ed S<strong>pe</strong>ed Flow Darcy gradient<br />

Pi<strong>pe</strong> size thickness for new of flow Flow of flow US Reynolds friction (m/m)<br />

(inch) (inch) pi<strong>pe</strong> (m/s) L/s (ft/s) gpm number factor or (ft/ft)<br />

4� Sch<br />

160<br />

5� Sch 40<br />

5� Sch 80<br />

5� Sch<br />

160<br />

6� Sch 40<br />

6� Sch 80<br />

6� Sch<br />

160<br />

8� Sch 40<br />

8� Sch 80<br />

Steel 0.531<br />

Rubber 0.250<br />

Steel 0.257<br />

Rubber 0.250<br />

Steel 0.375<br />

Rubber 0.250<br />

Steel 0.625<br />

Rubber 0.250<br />

Steel 0.280<br />

Rubber 0.250<br />

Steel 0.432<br />

Rubber 0.250<br />

Steel 0.719<br />

Rubber 0.250<br />

Steel 0.322<br />

Rubber 0.375<br />

Steel 0.500<br />

Rubber 0.375<br />

0.002015<br />

0.001281<br />

0.001349<br />

0.00152<br />

0.001063<br />

0.001124<br />

0.001262<br />

0.000817<br />

0.000859<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

04.4<br />

8.7<br />

13.1<br />

17.4<br />

21.8<br />

10.8<br />

21.5<br />

32.3<br />

43.1<br />

53.8<br />

9.7<br />

19.4<br />

29.1<br />

38.8<br />

48.5<br />

7.61<br />

15.2<br />

22.8<br />

30.4<br />

38.0<br />

15.6<br />

31.3<br />

47<br />

62.5<br />

78.2<br />

14<br />

28<br />

42<br />

56<br />

70<br />

11.1<br />

22.2<br />

33.3<br />

44.4<br />

55.5<br />

26.5<br />

53<br />

79.5<br />

106<br />

132<br />

24<br />

48<br />

72<br />

03.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

069<br />

138<br />

207<br />

276<br />

345<br />

171<br />

341<br />

512<br />

683<br />

853<br />

154<br />

308<br />

461<br />

615<br />

769<br />

121<br />

241<br />

362<br />

482<br />

603<br />

248<br />

496<br />

744<br />

992<br />

1240<br />

222<br />

443<br />

665<br />

886<br />

1108<br />

176<br />

352<br />

528<br />

703<br />

879<br />

420<br />

840<br />

1,260<br />

1,680<br />

2,100<br />

380<br />

760<br />

1,139<br />

074,422<br />

148,844<br />

223,266<br />

297,688<br />

372,110<br />

117,069<br />

234,137<br />

351,206<br />

468,274<br />

585,343<br />

111,125<br />

222,250<br />

333,375<br />

444,500<br />

555,625<br />

98,425<br />

196,850<br />

295,275<br />

393,700<br />

492,125<br />

141,148<br />

282,296<br />

423,443<br />

564,591<br />

705,739<br />

133,426<br />

266,852<br />

400,279<br />

533,705<br />

667,131<br />

118.847<br />

237,693<br />

356,540<br />

475,386<br />

594,233<br />

183,667<br />

367,335<br />

551,002<br />

734,670<br />

918,337<br />

174,625<br />

349,250<br />

523,875<br />

0.0259<br />

0.0248<br />

0.0244<br />

0.0242<br />

0.0241<br />

0.023<br />

0.0221<br />

0.0217<br />

0.0215<br />

0.0214<br />

0.0233<br />

0.0224<br />

0.022<br />

0.0218<br />

0.0217<br />

0.024<br />

0.0231<br />

0.0227<br />

0.0225<br />

0.0224<br />

0.0219<br />

0.0211<br />

0.0207<br />

0.0206<br />

0.0204<br />

0.0222<br />

0.0214<br />

0.0210<br />

0.0208<br />

0.0207<br />

0.0229<br />

0.0219<br />

0.0215<br />

0.0215<br />

0.0213<br />

0.0205<br />

0.0198<br />

0.0195<br />

0.0193<br />

0.0192<br />

0.0208<br />

0.0199<br />

0.0197<br />

0.0178<br />

0.068<br />

0.151<br />

0.265<br />

0.412<br />

0.01<br />

0.038<br />

0.085<br />

0.150<br />

0.233<br />

0.0107<br />

0.0410<br />

0.0901<br />

0.160<br />

0.249<br />

0.0124<br />

0.0478<br />

0.1058<br />

0.1865<br />

0.2898<br />

248<br />

496<br />

744<br />

992<br />

1240<br />

0.0085<br />

0.0326<br />

0.0723<br />

0.1274<br />

0.198<br />

0.0098<br />

0.0377<br />

0.0835<br />

0.1472<br />

0.2288<br />

0.0057<br />

0.0219<br />

0.0486<br />

0.0856<br />

0.1331<br />

0.006<br />

0.0233<br />

0.0517


TABLE 2-10 Continued<br />

FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

2.31<br />

Relative Energy<br />

Wall roughness S<strong>pe</strong>ed S<strong>pe</strong>ed Flow Darcy gradient<br />

Pi<strong>pe</strong> size thickness for new of flow Flow of flow US Reynolds friction (m/m)<br />

(inch) (inch) pi<strong>pe</strong> (m/s) L/s (ft/s) gpm number factor or (ft/ft)<br />

8� Sch 80<br />

8� Sch<br />

160<br />

10� Sch<br />

40 S<br />

10� Sch<br />

60 X<br />

10� Sch<br />

120 XX<br />

12� S<br />

12� X<br />

12� Sch<br />

120 XX<br />

14� S<br />

Steel 0.906<br />

Rubber 0.375<br />

Steel 0.365<br />

Rubber 0.375<br />

Steel 0.500<br />

Rubber 0.375<br />

Steel 1.000<br />

Rubber 0.375<br />

Steel 0.375<br />

Rubber 0.375<br />

Steel 0.500<br />

Rubber 0.375<br />

Steel 1.000<br />

Rubber 0.375<br />

Steel 0.375<br />

Rubber 0.375<br />

0.000974<br />

0.000637<br />

0.000656<br />

0.000738<br />

0.000525<br />

0.000537<br />

0.000591<br />

0.000472<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

96<br />

120<br />

19<br />

37<br />

6<br />

75<br />

93<br />

43.5<br />

87<br />

131<br />

174<br />

218<br />

41<br />

82<br />

123<br />

164<br />

205<br />

32<br />

65<br />

97<br />

130<br />

162<br />

64.1<br />

128.3<br />

192.4<br />

256.5<br />

320.7<br />

61<br />

123<br />

184<br />

245<br />

307<br />

51<br />

101<br />

152<br />

203<br />

253<br />

79<br />

158<br />

238<br />

317<br />

396<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

1,519<br />

1898<br />

295<br />

590<br />

886<br />

1,181<br />

1,476<br />

690<br />

1,380<br />

2,071<br />

2,761<br />

3,451<br />

651<br />

1,301<br />

1,952<br />

2,602<br />

3,253<br />

514<br />

1,028<br />

1,542<br />

2,056<br />

2,570<br />

1,017<br />

2,033<br />

3,049<br />

4,066<br />

5,082<br />

972<br />

1,943<br />

2,915<br />

3,887<br />

4,859<br />

803<br />

1606<br />

2409<br />

3212<br />

4015<br />

1,255<br />

2,510<br />

3,765<br />

5,020<br />

6,275<br />

698,500<br />

873,125<br />

154,000<br />

308,000<br />

462,001<br />

616,001<br />

770,001<br />

235,458<br />

470,916<br />

706,374<br />

941,832<br />

1,177,290<br />

228,600<br />

457,200<br />

658,800<br />

914,400<br />

1,143,000<br />

203,200<br />

406,400<br />

609,600<br />

812,800<br />

1,016,000<br />

285,750<br />

571,500<br />

857,250<br />

1,143,000<br />

1,428,750<br />

279,400<br />

558,800<br />

838,200<br />

1,117,600<br />

1,397,000<br />

254,000<br />

508,000<br />

762,000<br />

1,016,000<br />

1,270,000<br />

317,500<br />

635,000<br />

952,500<br />

1,270,000<br />

1,587,500<br />

0.0195<br />

0.0194<br />

0.0215<br />

0.0206<br />

0.0203<br />

0.0201<br />

0.020<br />

0.0194<br />

0.0186<br />

0.0183<br />

0.0182<br />

0.0181<br />

0.0195<br />

0.0188<br />

0.0185<br />

0.0183<br />

0.0182<br />

0.0201<br />

0.0198<br />

0.0189<br />

0.0188<br />

0.0187<br />

0.01853<br />

0.0178<br />

0.01755<br />

0.0174<br />

0.0173<br />

0.0186<br />

0.0179<br />

0.0176<br />

0.0175<br />

0.0174<br />

0.0190<br />

0.0183<br />

0.0180<br />

0.0179<br />

0.0179<br />

0.0181<br />

0.0174<br />

0.0171<br />

0.0170<br />

0.0169<br />

0.0912<br />

0.142<br />

0.0071<br />

0.0273<br />

0.0604<br />

0.1065<br />

0.1656<br />

0.0042<br />

0.0161<br />

0.0357<br />

0.063<br />

0.098<br />

0.0044<br />

0.0167<br />

0.0371<br />

0.0653<br />

0.1016<br />

0.0050<br />

0.0193<br />

0.0429<br />

0.0756<br />

0.1174<br />

0.0033<br />

0.0127<br />

0.0282<br />

0.0497<br />

0.0722<br />

0.0034<br />

0.0131<br />

0.029<br />

0.051<br />

0.079<br />

0.038<br />

0.0147<br />

0.0326<br />

0.0574<br />

0.0892<br />

0.0029<br />

0.0112<br />

0.0248<br />

0.0437<br />

0.0679<br />

(continued)


2.32 CHAPTER TWO<br />

TABLE 2-10 Continued<br />

Relative Energy<br />

Wall roughness S<strong>pe</strong>ed S<strong>pe</strong>ed Flow Darcy gradient<br />

Pi<strong>pe</strong> size thickness for new of flow Flow of flow US Reynolds friction (m/m)<br />

(inch) (inch) pi<strong>pe</strong> (m/s) L/s (ft/s) gpm number factor or (ft/ft)<br />

14� X<br />

14� Sch<br />

120<br />

16� Sch<br />

30 S<br />

16� X<br />

16� Sch<br />

120<br />

18� S<br />

18� X<br />

18� Sch<br />

120<br />

20� Sch<br />

20S<br />

Steel 0.500<br />

Rubber 0.375<br />

Steel 1.062<br />

Rubber 0.375<br />

Steel 0.375<br />

Rubber 0.375<br />

Steel 0.500<br />

Rubber 0.375<br />

Steel 1.291<br />

Rubber 0.375<br />

Steel 0.375<br />

Rubber 0.375<br />

Steel 0.500<br />

Rubber 0.375<br />

Steel 1.375<br />

Rubber 0.375<br />

Steel 0.375<br />

Rubber 0.375<br />

0.000482<br />

0.000531<br />

0.000407<br />

0.000414<br />

0.000466<br />

0.000358<br />

0.000363<br />

0.000407<br />

0.000319<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

76<br />

152<br />

228<br />

304<br />

380<br />

63<br />

125<br />

188<br />

251<br />

314<br />

107<br />

213<br />

319<br />

426<br />

532<br />

103<br />

206<br />

309<br />

412<br />

515<br />

81<br />

162<br />

243<br />

324<br />

405<br />

138<br />

276<br />

414<br />

552<br />

690<br />

134<br />

268<br />

401<br />

535<br />

669<br />

107<br />

213<br />

320<br />

426<br />

533<br />

173<br />

346<br />

520<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

1,205<br />

2,410<br />

3,616<br />

4,820<br />

6,026<br />

997<br />

1988<br />

2983<br />

3977<br />

4971<br />

1,689<br />

3,377<br />

5,066<br />

6,755<br />

8,443<br />

1,631<br />

3,262<br />

4,893<br />

6,524<br />

8,155<br />

1,289<br />

2,578<br />

3867<br />

5155<br />

6444<br />

2,187<br />

4,373<br />

6,560<br />

8,746<br />

10,933<br />

2,121<br />

4,241<br />

6,362<br />

8,483<br />

10,604<br />

1,689<br />

3,377<br />

5,066<br />

6,755<br />

8,443<br />

2,7495<br />

497<br />

8,246<br />

311,150<br />

622,300<br />

933,450<br />

1,244,600<br />

1,555,750<br />

282,600<br />

565,201<br />

847,801<br />

1,130,402<br />

1,413,002<br />

368,300<br />

736,600<br />

1,104,900<br />

1,473,200<br />

1,841,500<br />

361,950<br />

713,900<br />

1,085,850<br />

1,447,800<br />

1,809,750<br />

321,767<br />

643,534<br />

965,302<br />

1,287,069<br />

1,608,836<br />

419,100<br />

838,200<br />

1,257,300<br />

1,676,400<br />

2,095,500<br />

412,750<br />

825,500<br />

1,238,250<br />

1,651,000<br />

2,063,750<br />

368,300<br />

736,600<br />

1,104,900<br />

1,473,200<br />

1,841,500<br />

469,900<br />

939,800<br />

1,409,700<br />

0.0182<br />

0.0175<br />

0.0172<br />

0.0171<br />

0.0170<br />

0.0186<br />

0.0179<br />

0.0176<br />

0.0175<br />

0.0174<br />

0.0175<br />

0.0168<br />

0.0166<br />

0.0165<br />

0.0164<br />

0.0176<br />

0.0169<br />

0.0167<br />

0.0165<br />

0.0164<br />

0.0180<br />

0.0174<br />

0.0171<br />

0.0169<br />

0.0169<br />

0.0170<br />

0.0164<br />

0.0161<br />

0.0160<br />

0.0159<br />

0.0171<br />

0.0164<br />

0.0162<br />

0.0160<br />

0.0159<br />

0.0175<br />

0.0168<br />

0.0166<br />

0.0165<br />

0.0164<br />

0.0166<br />

0.0159<br />

0.0157<br />

0.003<br />

0.0115<br />

0.0254<br />

0.0447<br />

0.0695<br />

0.0034<br />

0.0129<br />

0.0286<br />

0.0503<br />

0.0783<br />

0.0024<br />

0.0093<br />

0.0207<br />

0.0364<br />

0.0566<br />

0.0025<br />

0.0095<br />

0.0211<br />

0.0372<br />

0.0578<br />

0.0029<br />

0.011<br />

0.0244<br />

0.0429<br />

0.0668<br />

0.002<br />

0.008<br />

0.0176<br />

0.0311<br />

0.0484<br />

0.0021<br />

0.0081<br />

0.0180<br />

0.0317<br />

0.0493<br />

0.0024<br />

0.0093<br />

0.0207<br />

0.0364<br />

0.0566<br />

0.0018<br />

0.0069<br />

0.0154


TABLE 2-10 Continued<br />

2-5 DYNAMICS OF THE BOUNDARY LAYER<br />

Boundary layer theory has been extensively covered by a number of authors. A book by<br />

Schlichting (1968) is considered one of the classical references on this subject. When a<br />

uniform flow approaches a plate, the particles at the wall of the plate are slowed down by<br />

the dynamic viscosity of the fluid. A layer called the boundary layer develops. When the<br />

flow enters a pi<strong>pe</strong>, effects develop at the entrance until the flow is uniform.<br />

2-5-1 Entrance Length<br />

FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

2.33<br />

Relative Energy<br />

Wall roughness S<strong>pe</strong>ed S<strong>pe</strong>ed Flow Darcy gradient<br />

Pi<strong>pe</strong> size thickness for new of flow Flow of flow US Reynolds friction (m/m)<br />

(inch) (inch) pi<strong>pe</strong> (m/s) L/s (ft/s) gpm number factor or (ft/ft)<br />

20� Sch<br />

20S<br />

20� Sch<br />

30 X<br />

24� Sch<br />

20 S<br />

24� Sch S<br />

Steel 0.500<br />

Rubber 0.375<br />

Steel 0.375<br />

Rubber 0.375<br />

Steel 0.500<br />

Rubber 0.375<br />

0.000324<br />

0.000262<br />

0.000265<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

694<br />

867<br />

169<br />

338<br />

506<br />

675<br />

844<br />

257<br />

513<br />

780<br />

1,026<br />

1,283<br />

251<br />

502<br />

753<br />

1,003<br />

1,254<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

10,995<br />

13,744<br />

2,675<br />

5,350<br />

8,025<br />

10,670<br />

13,375<br />

4,066<br />

8,132<br />

12,198<br />

16,284<br />

20,330<br />

3,976<br />

7,952<br />

11,930<br />

15,904<br />

19,880<br />

1,879,600<br />

2,349,500<br />

463,550<br />

927,100<br />

1,390,650<br />

1,854,200<br />

2,317,750<br />

571,500<br />

1,143,000<br />

1,714,500<br />

2,286,000<br />

2,857,500<br />

565,150<br />

1,130,300<br />

1,695,450<br />

2,260,600<br />

2,825,700<br />

0.0156<br />

0.0155<br />

0.0166<br />

0.0160<br />

0.0158<br />

0.0156<br />

0.0155<br />

0.0159<br />

0.0153<br />

0.0151<br />

0.0150<br />

0.0149<br />

0.0159<br />

0.0154<br />

0.0151<br />

0.0150<br />

0.0149<br />

0.0271<br />

0.0421<br />

0.0018<br />

0.0070<br />

0.0156<br />

0.0275<br />

0.0428<br />

0.0014<br />

0.0055<br />

0.0121<br />

0.0214<br />

0.0332<br />

0.0014<br />

0.0055<br />

0.0123<br />

0.0216<br />

0.0336<br />

*Dimensions are based on ANSI B36.1 and B31.1. Absolute roughness of rubber is input as 150 �m<br />

(0.000492 ft).<br />

Flow in a pi<strong>pe</strong> at relatively low s<strong>pe</strong>ed when the Reynolds number is smaller than 2500 is<br />

characterized by a certain distance called “entrance length,” over which the velocity profile<br />

takes the final parabolic sha<strong>pe</strong> shown in Figure 2-14. The length Le is expressed as<br />

Le = 0.028 DiRe For turbulent flows, the entrance length is equivalent to 50 times the inner diameter.


2.34 CHAPTER TWO<br />

rs (0.000492 ft)<br />

Hydraulic friction gradient (m/m)<br />

Hydraulic friction gradient (ft/ft)<br />

1.0<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

sch 160<br />

1 m/s<br />

2 m/s<br />

sch 80<br />

sch 40<br />

3 m/s<br />

4 m/s<br />

0.0 2.0 4.0 6.0<br />

Flow Rate (L/s)<br />

1.0<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

sch 160<br />

6.6 ft/sec<br />

3.3 ft/sec<br />

5 m/s<br />

8.0<br />

rometers (0.000492 ft)<br />

sch 80<br />

sch 40<br />

9.9ft/sec<br />

13.2 ft/sec<br />

16.5 ft/sec<br />

0.0 20 40 60 80 100 120<br />

Flow Rate (US gallons/min)<br />

FIGURE 2-5 Hydraulic friction gradient for water in a rubber-lined 2� pi<strong>pe</strong>, Sch 40, Sch 80<br />

and Sch 160. Caculations for 2� steel pi<strong>pe</strong>. Rubber thickness = 0.250� (6.4 mm). Rubber roughness<br />

= 150 �m (0.000492 ft).


Hydraulic friction (ft/ft) gradient (ft/ft) Hydraulic friction (m/m) gradient (m/m)<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

1 m/s<br />

2 m/s<br />

FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

3 m/s<br />

0.0 2.0 4.0 6.0<br />

0.1<br />

3.3 ft/sec<br />

0.0<br />

0<br />

6.6 ft/sec<br />

9.9ft/sec<br />

4 m/s<br />

8.0<br />

13.2 ft/sec<br />

2 m/s<br />

sch 160<br />

5 m/s<br />

2.35<br />

2-5-2 Friction Velocity<br />

Prandtl proposed a concept of friction velocity:<br />

Uf = = U (2-25)<br />

�� ��<br />

Blasius conducted tests on turbulent flows in pi<strong>pe</strong>s and develo<strong>pe</strong>d an equation for the<br />

shear wall stress in terms of the maximum velocity outside the boundary layer.<br />

�w = 0.0225 �(Umax) 7/4� � 1/4<br />

(2-26)<br />

The local magnitude of the velocity in a boundary layer at a height y above the wall is<br />

= 8.73(y + ) n (2-27)<br />

where n = 1/7 to 1/9 and where the nondimensional height parameter y + �w fN � �<br />

� 2<br />

r<br />

�<br />

R<br />

u<br />

�<br />

Uf<br />

is defined as the<br />

relative distance from the wall:<br />

sch 80<br />

sch 40<br />

10.0 12.0 14.0 16.0 18.0 20.0<br />

Flow Rate (L/s)<br />

16.5 ft/sec<br />

sch 160<br />

6.6 ft/sec<br />

3 m/s<br />

sch 80<br />

9.9 ft/sec<br />

sch 40<br />

4 m/s<br />

Rubber Lined<br />

Steel Pi<strong>pe</strong>s<br />

20 40 60 80 100 120 140 160 180 200 220 240 260 220<br />

Flow Rate (US gallons/min)<br />

HDPE SDR 11<br />

5 m/s<br />

Rubber Lined<br />

Steel Pi<strong>pe</strong>s<br />

HDPE SDR 11<br />

16.5 ft/sec<br />

13.2 ft/sec<br />

FIGURE 2-6 Hydraulic friction gradient for water in a rubber-lined 3� pi<strong>pe</strong>, Sch 40, Sch 80,<br />

and Sch 160. Rubber thickness = 0.25� (6.4 mm). Rubber roughness = 150 �m (0.000492 ft).<br />

2-HDPE line 3� SDR11 roughness 1.5 �m (0.00000492 ft).<br />

240<br />

260


2.36<br />

Energy gradient (m/m)<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

2 m/s<br />

1 m/s<br />

0.0<br />

0.0<br />

4 m/s<br />

sch 160<br />

sch 80<br />

sch 40<br />

5 m/s<br />

3 m/s 5 m/s<br />

2 m/s<br />

Rubber Lined<br />

Steel Pi<strong>pe</strong>s<br />

3 m/s<br />

4 m/s<br />

10 20 30 40<br />

Flow Rate (L/s)<br />

HDPE<br />

SDR<br />

11<br />

Energy gradient (ft/ft)<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

0.0<br />

9.9 ft/sec<br />

sch 160<br />

13.2 ft/sec<br />

Rubber Lined<br />

Steel Pi<strong>pe</strong>s<br />

sch 80<br />

sch 40<br />

9.9 ft/sec<br />

6.6 ft/sec<br />

3.3 ft/sec<br />

100 200 300 400 500<br />

Flow Rate (US gallons/min)<br />

16.5 ft/sec<br />

6.6 ft/sec 13.2 ft/sec<br />

16.5 ft/sec<br />

HDPE<br />

SDR<br />

11<br />

FIGURE 2-7 Hydraulic friction gradient for water in a rubber-lined 4� pi<strong>pe</strong>, Sch 40, Sch 80, and Sch 160. Rubber thickness = 0.25� (6.4<br />

mm). Rubber roughness = 150 �m (0.000492 ft). 2-HDPE line 3� SDR11 roughness 1.5 �m (0.00000492 ft).


Energy gradient (ft/ft)<br />

Energy gradient (m/m)<br />

0.30<br />

0.25<br />

0.20<br />

0.15<br />

0.10<br />

0.05<br />

0.0<br />

0.30<br />

0.25<br />

0.20<br />

0.15<br />

0.10<br />

0.05<br />

0.0<br />

0.0<br />

0.0<br />

1 m/s<br />

FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

2 m/s<br />

3 m/s<br />

4 m/s<br />

Rubber Lined Steel<br />

5 m/s<br />

sch 160<br />

2.37<br />

y + = (2-28)<br />

and where u = velocity of the flow at distance y.<br />

The boundary layer can be divided into a number of sections. At small values of y + up<br />

to 5, the velocity profile is linear in a sublayer (see Figure 2-15). The flow is considered<br />

to be laminar in the sublayer. Above y + = 5, a buffer zone develops up to y + Uf� �<br />

�<br />

= 50 and turbulence<br />

develops. The thickness of the boundary viscous sublayer � is usually expressed<br />

as<br />

11.6 � 11.6 � fD � = � = � � (2-29)<br />

�(���w�)� �U �� 8<br />

In the turbulent region, the velocity profile is established as<br />

u<br />

Ufy � = 5.75 log10��� + 5.5 (2-30)<br />

Uf<br />

�<br />

sch 80<br />

Flow Rate (L/s)<br />

sch 40<br />

10 20 30 40 50 60 70 80<br />

9.9 ft/sec<br />

6.6 ft/sec<br />

3.3 ft/sec<br />

13.2 ft/sec<br />

sch 160<br />

200 400 600 800 1000 1200<br />

Flow Rate ( US gallons/sec)<br />

16.5 ft/sec<br />

HDPE SDR11<br />

FIGURE 2-8 Hydraulic friction gradient for water in a rubber-lined 6� pi<strong>pe</strong>, Sch 40, Sch 80,<br />

and Sch 160. Rubber thickness = 0.25� (6.4 mm). Rubber roughness = 150 �m (0.000492 ft).<br />

2-HDPE line 3� SDR11 roughness 1.5 �m (0.00000492 ft).<br />

sch 80<br />

sch 40<br />

HDPE SDR11


2.38<br />

Hydraulic friction gradient (m/m)<br />

0.18<br />

0.16<br />

0.14<br />

0.12<br />

0.10<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

0.0<br />

2 m/s<br />

1 m/s<br />

3 m/s<br />

0.0 20 40 60<br />

4 m/s<br />

sch 160<br />

80<br />

5 m/s<br />

sch 80<br />

sch 40<br />

Rubber Lined Steel<br />

0.18<br />

0.16<br />

0.14<br />

Hydraulic friction gradient (ft/ft)<br />

0.12<br />

0.10<br />

0.08<br />

13.2 ft/sec<br />

0.06<br />

9.9 ft/sec<br />

HDPE SDR11 0.04<br />

HDPE SDR11<br />

0.02<br />

6.6 ft/sec<br />

0.0<br />

3.3 ft/sec<br />

100 120 140<br />

0 400 800 1200 1600 2000<br />

Flow Rate (L/s)<br />

Flow Rate (US gallons/min)<br />

sch 160 16.5 ft/sec<br />

sch 80<br />

sch 40<br />

Rubber Lined Steel<br />

FIGURE 2-9 Hydraulic friction gradient for water in a rubber-lined 8� pi<strong>pe</strong>, Sch 40, Sch 80, and Sch 160. Rubber thickness = 0.375� (9.5 mm). Rubber<br />

roughness = 150 �m (0.000492 ft). 2-HDPE roughness 1.5 �m (0.00000492 ft).


3.3 ft/sec<br />

Hydraulic friction gradient (m/m)<br />

Hydraulic friction gradient (ft/ft)<br />

0.12<br />

0.10<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

0.0<br />

0.12<br />

0.10<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

0.0<br />

0.0<br />

FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

2 m/s<br />

1 m/s<br />

40<br />

80<br />

3 m/s<br />

9.9 ft/sec<br />

6.6 ft/sec<br />

10" sch 40<br />

1 m/s<br />

120<br />

10" sch 40<br />

2 m/s<br />

160<br />

5 m/s<br />

16.5 ft/sec<br />

4 m/s<br />

4 m/s<br />

3 m/s<br />

12" sch 40<br />

200 240 280<br />

Flow Rate (L/s)<br />

9.9 ft/sec<br />

320<br />

0.0 800 1600 2400 3200 4000 4800<br />

Flow Rate ( US gallons/sec)<br />

6.6 ft/sec<br />

13.2 ft/sec<br />

12" sch 40<br />

13.2 ft/sec<br />

5 m/s<br />

16.5 ft/sec<br />

2.39<br />

FIGURE 2-10 Hydraulic friction gradient for water in a rubber-lined 10� pi<strong>pe</strong>, Sch 40 and<br />

12�, Sch 40 pi<strong>pe</strong>s. Rubber thickness = 0.375� (9.5 mm). Rubber roughness = 150 �m<br />

(0.000492 ft). 2-HDPE roughness 1.5 �m (0.00000492 ft).


2.40 CHAPTER TWO<br />

Hydraulic friction Hydraulic friction<br />

gradient (ft/ft)<br />

gradient (m/m)<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

0.0<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

0.0<br />

0.0<br />

0.0<br />

The reader is encouraged to review the work of Schlichling (1968) for details of<br />

boundary layer theory.<br />

Example 2-5<br />

Determine the boundary viscous sublayer thickness and friction velocity of Example 2-4.<br />

Solution in SI Units<br />

In Example 2-4, the Darcy friction factor was determined to be 0.0178. In Section 2.31 it<br />

was stated that fD = 4fN, therefore<br />

0.0178<br />

fN = � = 0.00445<br />

4<br />

Since the velocity of the flow is 1.467 m/s,<br />

ff Uf = U�� �<br />

2<br />

10" SDR 11<br />

5 m/s<br />

12" SDR 11<br />

5 m/s<br />

40 80 120 160 200 240 280 320<br />

Flow Rate (L/s)<br />

10" SDR 11<br />

16.5 ft/sec<br />

16.5 ft/sec<br />

12" SDR 11<br />

800 1600 2400 3200 4000 4800<br />

Flow Rate ( US gallons/sec)<br />

FIGURE 2-11 Hydraulic friction gradient of water in 10� and 12� SDR11 HDPE pi<strong>pe</strong>s.<br />

Roughness 1.5 �m (0.00000492 ft).<br />

0.00445<br />

= 1.467��<br />

� = 0.0692 m/s<br />

2<br />

From Equation 2.28 the thickness of the viscous sublayer is calculated as<br />

� = �� f 11.6 � D<br />

� �8<br />

�U


Hydraulic friction gradient (m/m)<br />

Hydraulic friction gradient (ft/ft)<br />

0.09<br />

0.08<br />

0.07<br />

0.06<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0.0<br />

FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

2 m/s<br />

1 m/s<br />

3 m/s<br />

4 m/s<br />

14" S (Sch 30)<br />

� = �� = (4.72) 10–7 11.6 (0.00129) 0.0178<br />

�� � m<br />

1020 (1.467) 8<br />

Solution in USCS Units<br />

0.0178<br />

fN = � = 0.00445<br />

4<br />

Since the average velocity flow in 4.8 ft/s,<br />

0.00445<br />

Uf = 4.8 � 0.2264 ft/s<br />

��2 2.41<br />

FIGURE 2-12 Hydraulic friction gradient for water in rubber-lined 14� pi<strong>pe</strong>, Sch 40 and 16�<br />

S and 18� S pi<strong>pe</strong>s. Wall thickness = 0.375� (9.5 mm). Rubber thickness = 0.375� (9.5 mm).<br />

Rubber roughness = 150 �m (0.000492 ft).<br />

16" S<br />

18" S<br />

5 m/s<br />

0.0 100 200 300 400 500 600 700<br />

009<br />

Flow Rate (L/s)<br />

0.09<br />

0.08<br />

0.07<br />

0.06<br />

Flow Rate (L/s)<br />

0.05<br />

0.04<br />

13.2 ft/sec<br />

16.5 ft/sec<br />

0.03<br />

0.02<br />

0.01<br />

0.0<br />

9.9 ft/sec<br />

6.6 ft/sec<br />

3.3 ft/sec<br />

0.0 2000 4000 6000 8000 10000<br />

Flow Rate (US gallons/min)<br />

14" S (Sch 30)<br />

16" S<br />

18" S


2.42 CHAPTER TWO<br />

Hydraulic friction gradient (m/m)<br />

Hydraulic friction gradient (ft/ft)<br />

0.09<br />

0.08<br />

0.07<br />

0.06<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0.0<br />

0.09<br />

0.08<br />

0.07<br />

0.06<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0.0<br />

0.0 100 200 300<br />

0.0<br />

14" SDR 11<br />

14" SDR 11<br />

400<br />

5 m/s<br />

16" SDR 11<br />

5 m/s<br />

500 600 700<br />

Flow Rate (L/s)<br />

16" SDR 11<br />

2000 4000 6000 8000 10000<br />

Flow Rate (US gallons/min)<br />

g 2-<br />

18" SDR 11<br />

5 m/s<br />

18" SDR 11<br />

16.5 ft/sec<br />

FIGURE 2-13 Hydraulic friction gradient of water in 14�, 16� and 18� SDR 11 HDPE pi<strong>pe</strong>s.<br />

Absolute roughness 1.5 �m (0.00000492 ft).


FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

2.43<br />

TABLE 2-11 Flow and Hydraulic Friction Gradient for HDPE Pi<strong>pe</strong>s SDR11 at a S<strong>pe</strong>ed<br />

of 1 m/s to 5 m/s (3.28 to 16.4 ft/sec) for Water at a Density of 1000 kg/m 3 (62.3 lbs/ft 3 ) and<br />

a Kinematic Viscosity of 1 cP (2.09 × 10 –5 lbf-sec/ft 2 ) Using the Swamee–Jain Equation<br />

Pi<strong>pe</strong> Friction<br />

inner S<strong>pe</strong>ed S<strong>pe</strong>ed Flow Darcy gradient<br />

Pi<strong>pe</strong> size diameter Relative of flow Flow of flow US Reynolds friction (m/m)<br />

(in) (in) roughness (m/s) L/s (ft/s) gpm number factor or (ft/ft)<br />

3<br />

4<br />

5<br />

6<br />

7�<br />

8�<br />

10�<br />

12�<br />

2.826<br />

3.633<br />

4.49<br />

5.349<br />

5.7510<br />

7.270�<br />

8.675�<br />

10.293�<br />

0.0000053<br />

00000041<br />

0.0000033<br />

0.0000028<br />

0.0000026<br />

0.0000022<br />

0.0000017<br />

0.0000015<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

4.0<br />

8.0<br />

12.1<br />

16.2<br />

20.2<br />

6.7<br />

13.4<br />

20.1<br />

26.8<br />

33.4<br />

10.2<br />

204<br />

306<br />

408<br />

510<br />

15<br />

29<br />

44<br />

58<br />

73<br />

17<br />

34<br />

51<br />

67<br />

84<br />

25<br />

49<br />

74<br />

98<br />

123<br />

38<br />

76<br />

114<br />

153<br />

191<br />

54<br />

107<br />

161<br />

215<br />

268<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

64<br />

128<br />

192<br />

257<br />

321<br />

106<br />

212<br />

318<br />

424<br />

530<br />

162<br />

324<br />

486<br />

648<br />

810<br />

230<br />

460<br />

689<br />

919<br />

1,150<br />

266<br />

531<br />

797<br />

1,063<br />

1,328<br />

389<br />

779<br />

1,168<br />

1,558<br />

1,947<br />

605<br />

1,210<br />

1,815<br />

2,420<br />

3,025<br />

851<br />

1,702<br />

2,552<br />

3,404<br />

4,254<br />

71,780<br />

143,561<br />

215,341<br />

287,122<br />

358,902<br />

92,278<br />

184,556<br />

276,835<br />

369,113<br />

461,391<br />

114,046<br />

228,092<br />

342,138<br />

456,184<br />

570,230<br />

135,865<br />

271,729<br />

407,594<br />

543,458<br />

679,323<br />

146,075<br />

292,151<br />

438,226<br />

584,302<br />

730,377<br />

176,860<br />

353,720<br />

530,581<br />

707,441<br />

884,301<br />

220,447<br />

440,893<br />

661,340<br />

881,786<br />

1,102,233<br />

261,442<br />

522,884<br />

784,327<br />

1,045,769<br />

1,307,211<br />

0.01917<br />

0.01659<br />

0.01532<br />

0.0145<br />

0.01391<br />

0.0182<br />

0.0158<br />

0.0146<br />

0.0138<br />

0.01329<br />

0.01738<br />

0.01514<br />

0.01403<br />

0.01331<br />

0.01279<br />

0.01677<br />

0.01465<br />

0.01359<br />

0.01290<br />

0.01241<br />

0.01653<br />

0.01445<br />

0.01341<br />

0.01274<br />

0.01225<br />

0.0159<br />

0.0139<br />

0.01296<br />

0.01232<br />

0.01186<br />

0.0152<br />

0.0134<br />

0.0125<br />

0.0119<br />

0.0114<br />

0.01475<br />

0.0130<br />

0.0121<br />

0.0115<br />

0.0111<br />

0.0136<br />

0.0471<br />

0.0979<br />

0.167<br />

0.247<br />

0.01<br />

0.035<br />

0.0726<br />

0.1223<br />

0.1835<br />

0.0078<br />

0.0271<br />

0.0564<br />

0.0592<br />

0.1429<br />

0.0063<br />

0.0220<br />

0.0459<br />

0.0774<br />

0.1164<br />

0.0058<br />

0.0202<br />

0.0421<br />

0.0711<br />

0.1069<br />

0.0046<br />

0.0161<br />

0.0336<br />

0.0568<br />

0.0854<br />

0.0035<br />

0.0124<br />

0.0259<br />

0.0439<br />

0.0660<br />

0.0029<br />

0.0101<br />

0.0212<br />

0.0359<br />

0.0541<br />

(continued)


2.44 CHAPTER TWO<br />

TABLE 2-11 Continued<br />

Pi<strong>pe</strong> Friction<br />

inner S<strong>pe</strong>ed S<strong>pe</strong>ed Flow Darcy gradient<br />

Pi<strong>pe</strong> size diameter Relative of flow Flow of flow US Reynolds friction (m/m)<br />

(in) (in) roughness (m/s) L/s (ft/s) gpm number factor or (ft/ft)<br />

13�<br />

14�<br />

16�<br />

18�<br />

20�<br />

10.797�<br />

11.301�<br />

12.915�<br />

14.532�<br />

16.146�<br />

0.0000014<br />

0.0000013<br />

0.0000012<br />

0.000001<br />

0.0000009<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

1<br />

2<br />

3<br />

4<br />

5<br />

59<br />

118<br />

177<br />

236<br />

295<br />

65<br />

129<br />

194<br />

259<br />

324<br />

85<br />

169<br />

253<br />

338<br />

423<br />

107<br />

214<br />

321<br />

428<br />

535<br />

132<br />

264<br />

396<br />

528<br />

660<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

3.3<br />

6.6<br />

9.9<br />

13.2<br />

16.5<br />

936<br />

1,872<br />

2,808<br />

3,745<br />

4,681<br />

1,026<br />

2,051<br />

3,077<br />

4,103<br />

5,129<br />

1,340<br />

2,679<br />

4,020<br />

5,359<br />

6,698<br />

1,696<br />

3,392<br />

5,088<br />

6,784<br />

8,480<br />

2,094<br />

4,187<br />

6,281<br />

8,375<br />

10,469<br />

From the Equation 2.28, the thickness of the viscous sub-layer is<br />

� = �� = 1.56 × 10–6 11.6 × 2.7 × 10 0.0178<br />

� ft<br />

8<br />

–5<br />

��<br />

1.973 × 4.8<br />

2-6 PRESSURE LOSSES DUE TO CONDUITS<br />

AND FITTINGS<br />

274,244<br />

548,488<br />

822,731<br />

1,096,975<br />

1,371,219<br />

287,045<br />

574,091<br />

861,136<br />

1,148,182<br />

1,435,227<br />

328,041<br />

656,082<br />

984,123<br />

1,312,164<br />

1,640,205<br />

369,113<br />

738,226<br />

1,107,338<br />

1,476,451<br />

1,845,564<br />

410,108<br />

820,217<br />

1,230,325<br />

1,640,434<br />

2,050,542<br />

0.0146<br />

0.0129<br />

0.0120<br />

0.0114<br />

0.0110<br />

0.0145<br />

0.0128<br />

0.0119<br />

0.0113<br />

0.0109<br />

0.0141<br />

0.0125<br />

0.0116<br />

0.0110<br />

0.0107<br />

0.0138<br />

0.0122<br />

0.0114<br />

0.0109<br />

0.0105<br />

0.0136<br />

0.0120<br />

0.0112<br />

0.0107<br />

0.0103<br />

0.0027<br />

0.0096<br />

0.0201<br />

0.0340<br />

0.0512<br />

0.0026<br />

0.0091<br />

0.0190<br />

0.0322<br />

0.0485<br />

0.0022<br />

0.0078<br />

0.0163<br />

0.0276<br />

0.0416<br />

0.0019<br />

0.0068<br />

0.0142<br />

0.0240<br />

0.0362<br />

0.0017<br />

0.0060<br />

0.0125<br />

0.0213<br />

0.0321<br />

The sizing of pumps is based on determining pressure losses between the starting point<br />

(A) and the final delivery point (B) (Figure 2-16). It is important to know that the static<br />

pressure at (A) includes atmospheric pressure or the pressure of any pressurizing gas, and


D<br />

i<br />

FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

2.45<br />

the pressure due to the height of liquid above the centerline of the first im<strong>pe</strong>ller or im<strong>pe</strong>ller<br />

at suction. It is also important to know that static pressure at (B) includes any pressurizing<br />

gas at (B) and the height of liquid at (B) above the centerline of the pump’s im<strong>pe</strong>ller.<br />

The additional pressure losses between (A) and (B) include the friction losses and<br />

pressure losses in all the pi<strong>pe</strong> fittings such as valves, elbows, expansions, contraction<br />

branches, and bypasses. Pressure is also lost at entry and exit as well. Such pressure losses<br />

are expressed in terms of the Darcy–Weisbach equation and in terms of pressure loss<br />

factors for each fitting.<br />

Total pressure loss due to friction:<br />

�U<br />

Hf = fD + �Kf (2-31)<br />

2<br />

�U Lj � �<br />

Dij 2<br />

2<br />

�<br />

2<br />

where<br />

L j = length of the conduit j<br />

K f = pressure loss of the fitting f<br />

+<br />

+<br />

L e<br />

FIGURE 2-14 Entrance length for flows in pi<strong>pe</strong>s.<br />

V<br />

V<br />

Turbulent layer<br />

Buffer layer<br />

Viscous sublayer<br />

FIGURE 2-15 Boundary layer of flow over a plate.


g 6<br />

2.46 CHAPTER TWO<br />

H<br />

1<br />

5 Pi<strong>pe</strong> Diameters<br />

FIGURE 2-16 Simplified pumping system between two tanks.<br />

TABLE 2-12 Equivalent Length of Valves for Friction Loss of Calculations for<br />

Single-Phase Turbulent Flow<br />

Equivalent<br />

length/diameter<br />

Minimum recommended<br />

s<strong>pe</strong>ed for full disc lift<br />

Fitting ratio m/s Ft/s<br />

Gate valves 8<br />

Globe valves 340<br />

Angle valves 55<br />

Ball valves 3<br />

Butterfly valves 16<br />

Plug valves—straightway 18<br />

Plug valves—3 way through flow 30<br />

Plug valve—branch flow 90<br />

Stop check valve—straight through 400 2.12 6.96<br />

Stop check valve—angular 90 deg 200 2.89 9.49<br />

Swing check valve 300 2.32 7.59<br />

Lift check valve 55 5.4 17.7<br />

Tilting disc check valve 5–15 1.16–3.08 3.80 to 10.13<br />

Foot valve with strainer—pop<strong>pe</strong>t disc 420 0.58 1.90<br />

Foot valve with strainer—hinged disc 75 1.35 4.43<br />

H<br />

2


FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

Loss Factor K<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

0 2 4<br />

6 8 10<br />

Ratio R/D<br />

(bend radius/pi<strong>pe</strong> diam)<br />

The differential head that a pump must deliver to pump a liquid between (A) and (B) is<br />

therefore<br />

TDH = (P B – P A)/�g + (Z B – Z A) + H fOB + H fOA<br />

where<br />

H fOA = the pressure losses due to conduits and fittings between the tank (A) and the<br />

pump, including entry loss<br />

H fOB = the losses due to conduits and fittings between the pump and tank (B), including<br />

exit losses<br />

Table 2-12 presents examples of loss coefficients for fittings. Some practical considerations<br />

limit the use of fittings in <strong>slurry</strong> circuits. For example, elbows should have a minimum<br />

radius of three pi<strong>pe</strong> diameters to avoid short turns (see Figure 2-17). Such an approach<br />

minimizes wear.<br />

Example 2-6<br />

The fluid of Example 2-4 is pum<strong>pe</strong>d at a flow rate of 3500 gpm through a 16 × 14 pump.<br />

The steel fittings include 14� × 18� reducer, an 18� knife gate valve, three long radius 90°<br />

elbows with a diameter to radius ratio of 3. The length of the pi<strong>pe</strong> is 355 ft. Determine the<br />

total dynamic head if the liquid level in the suction tank is 10 ft and the level in the discharge<br />

tank is 50 ft above the centerline of the im<strong>pe</strong>ller. Ignore the length of the suction<br />

pi<strong>pe</strong> as negligible.<br />

Solution in SI Units<br />

Net static head:<br />

(50 ft – 10 ft) 0.3048 = 12.2 m<br />

Dynamic head at the entry of the pump:<br />

Di = 15.25 in or 0.387 m<br />

Suction Area = 0.1178 m 2<br />

2.47<br />

FIGURE 2-17 Loss Factors for rough wall bends for pi<strong>pe</strong>s 1–7� (after Crane Technical Bulletin<br />

No 410).


2.48 CHAPTER TWO<br />

S<strong>pe</strong>ed = = 1.875 m/s<br />

Dynamic head at suction:<br />

1.875<br />

= = 0.18 m<br />

For a sharp entrance pi<strong>pe</strong>, the recommended K factor is 0.5. Friction losses at the entry<br />

to the pump are calculated as follows:<br />

0.5 × 0.18 = 0.09 m<br />

On the discharge of the pump for a 14 × 18 reducer, the loss factor is calculated from<br />

the area ratio as<br />

2<br />

U<br />

�<br />

2 × 9.81<br />

2<br />

0.221<br />

�<br />

0.1178<br />

�<br />

2g<br />

K = �1 – � 2<br />

= �1 – � 2<br />

� � = 0.1681<br />

2<br />

2<br />

d 2 17.25<br />

� For an18 ft full-bore gate valve, the loss factor K = 0.10<br />

� For the elbow r/D = 3, the loss factor K = 0.14<br />

� For the reentry pi<strong>pe</strong> L/D = 65<br />

Total friction losses:<br />

355 × 0.3048 + 65 × 0.438<br />

0.09 m + � ��� × 0.0178 + (0.1681 + 0.10 + 3 × 0.14)�<br />

0.438<br />

1.467 2<br />

× � = 0.774 m<br />

2 × 9.81<br />

13.25 2<br />

TDH = 0.774 m + 12.2 m = 12.97 m<br />

Solution in USCS Units<br />

Net static head = 50 ft – 10 ft = 40 ft<br />

S<strong>pe</strong>ed in suction pi<strong>pe</strong> is calculated as follows:<br />

ID = 15.25� = 1.27 ft<br />

d 1 2<br />

Flow Rate = 7.79 ft 3 /s<br />

Area = 1.267 ft 2<br />

Velocity = = 6.15 ft/s<br />

Dynamic head at suction:<br />

6.15<br />

= 0.587 ft<br />

The K factor for sharp entrance in 0.5. Loss during suction is 0.5 × 0.587 = 0.293 ft.<br />

On the discharge of the pump the K factor is determined as in the SI unit solution. Total<br />

friction losses are:<br />

2<br />

7.79<br />

�<br />

1.267<br />

�<br />

2 × 32.2


FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

0.293 ft + �0.0178� � + 0.1681 + 0.10 + 0.42�<br />

355 + 65 × 1.27<br />

�� � = 2.44 ft<br />

1.27<br />

2 × 32.2<br />

TDH = 40 + 2.44 = 42.44 ft<br />

2-7 ORIFICE PLATES, NOZZLES, AND<br />

VALVE HEAD LOSSES<br />

The flow through an orifice plate, a nozzle, or a valve is reduced from the ideal theoretical<br />

value by a discharge coefficient:<br />

Q = CdQideal (2-32)<br />

The ideal theoretical flow is considered the product of s<strong>pe</strong>ed and area at the o<strong>pe</strong>ning of<br />

the orifice, valve, or nozzle.<br />

The theoretical velocity through the orifice is calculated as<br />

2�P<br />

Vth = �2�g�h� = � (2-33)<br />

�� R<br />

However, the velocity is typically smaller than the theoretical velocity:<br />

V o = C veV th<br />

where C ve = velocity coefficient.<br />

The flow through an o<strong>pe</strong>ning contracts from the full area. This is known as the vena<br />

contracta effect.<br />

Reentrant Sharp<br />

tube Edged<br />

V V<br />

Square<br />

Edged<br />

Length = 1/2 to<br />

Stream clears<br />

1 diameter sides<br />

4.8 2<br />

Reentrant<br />

tube<br />

C =0.52<br />

C =0.61 C =0.61 C =0.73<br />

d d d<br />

d<br />

V<br />

2.49<br />

Length = 2-1/2<br />

diameters<br />

FIGURE 2-18 Discharge coefficients of orifice plates and nozzles.


2.50 CHAPTER TWO<br />

For a thin plate or sharp-edged orifice, the vena contracta is assumed to be one half of<br />

an orifice diameter d1 downstream from the orifice, but in reality the distance may be from<br />

30% to 80% of d1. For flow of water at a high Reynolds number through a small orifice diameter,<br />

Lindeburg (1998) reported that the contracted area is approximately 61% to 63%.<br />

The coefficient of contraction is defined as<br />

area of vena contracta<br />

Cc = ���<br />

(2-34).<br />

orifice area<br />

The total discharge is the product of the reduced velocity and the contracted area:<br />

Q = C veV th(C c A 1)<br />

The product of the velocity coefficient by the area contraction coefficient is called the discharge<br />

coefficient:<br />

C d = C veC c<br />

(2-35)<br />

Q = C d A 1�(2�g�h�)� (2-36a)<br />

Q = Cd A1�(2���P�/��)� (2-36b)<br />

actual discharge<br />

Cd = ���<br />

theoretical discharge<br />

Typical values for discharge coefficients from nozzles and orifices are shown in Figure<br />

2-18. The Cameron Hydraulic Handbook (1977) recommends a further correction for<br />

large o<strong>pe</strong>nings when d2/d1 > 0.30:<br />

2gh<br />

Q = Cd A �� �� 1 – (d1/d2) 4<br />

(2-37)<br />

This equation works for liquids with a dynamic viscosity similar to the viscosity of water.<br />

The discharge vena contracta and velocity coefficient presented in Figure 2-18 are<br />

based on controlled flow conditions upstream. Flow disturbances can affect the magnitude<br />

of these coefficients.<br />

Manufacturers of valves in North America have develo<strong>pe</strong>d a valve coefficient to relate<br />

flow rate to pressure drop as Cv, which is defined as:<br />

�Ppsi Qgpm = Cv�� � (2-38)<br />

S.G.<br />

This coefficient is not dimensionally homogeneous and is not equal to the discharge<br />

coefficient from orifices and nozzles. Although the flow coefficient Cv was develo<strong>pe</strong>d for<br />

control valves, a relationship is often established for other fittings in terms of the K factor:<br />

(29.9)(d in) 2<br />

Cv = ��<br />

(2-39)<br />

�K�<br />

The reader should be very careful not to confuse C v (the flow coefficient commonly<br />

used in North America) with the discharge coefficient C d more commonly used in the rest<br />

of the world. Such a mix-up can lead to serious errors. C v is not used outside North America<br />

and has no relationship to the terms defined in Equations 2-34 to 2-37. The reader<br />

should avoid the common confusion that it sometimes creates.


FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

2.51<br />

FIGURE 2-19 Cross-section of a Series 39 <strong>slurry</strong> check valve. (Courtesy of Red Valve<br />

Company, Carnegie, PA, U.S.A.)<br />

FIGURE 2-20 Front view of a Series 39 <strong>slurry</strong> check valve. (Courtesy of Red Valve Company,<br />

Carnegie, PA, U.S.A.)


2.52 CHAPTER TWO<br />

FIGURE 2-21 Slurry knife-gate valve cross-sectional drawing. (Courtesy of Red Valve<br />

Company, Carnegie, PA, U.S.A.)<br />

FIGURE 2-22 Slurry knife-gate valve. (Courtesy of Red Valve Company, Carnegie, PA,<br />

U.S.A.)


Manufacturers of <strong>slurry</strong> valves have develo<strong>pe</strong>d very s<strong>pe</strong>cific designs to meet the requirements<br />

of wear and o<strong>pe</strong>ration without plugging. These include:<br />

� Rubber-lined check valves<br />

� Rubber-lined knife-gate valves<br />

� Rubber-lined pinch valves<br />

� Ceramic ball valves<br />

� Plug valves<br />

FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

2.53<br />

FIGURE 2-23 Slurry pinch valve, showing cut through the rubber sleeve. (Courtesy of Red<br />

Valve Company, Carnegie, PA, U.S.A.)<br />

S<strong>pe</strong>cial check valves are available for sewage and <strong>slurry</strong> flows. The Red Valve Company<br />

Series 39 valves (Figures 2-19 and 2-20) feature a s<strong>pe</strong>cial reinforced elastomer check<br />

sleeve. The valve check sleeve seals under reverse flow or back-pressure and o<strong>pe</strong>ns under<br />

pressure from the pump. It does not incorporate any discs that may wear on contact with<br />

<strong>slurry</strong>. This ty<strong>pe</strong> of valve is therefore different in design than the ty<strong>pe</strong> shown in books on<br />

water flows. The consultant engineer should therefore request from the manufacturer of<br />

the <strong>slurry</strong> check valves the estimated K factor for pressure losses. The Red Valve Company<br />

Series 39 <strong>slurry</strong> check valves are available in sizes up to 48� (1220 mm), with a choice<br />

of elastomers such as pure gum rubber, neoprene, Hypalon, chlorobutyl, Buna-N, EPDM,<br />

and Viton.<br />

Knife-gate valves for <strong>slurry</strong> flows (Figures 2-21 and 2-22) feature a metal gate sand-


2.54 CHAPTER TWO<br />

FIGURE 2-24 Principles of o<strong>pe</strong>ration of a pinch valve, pinched by a roller. (Courtesy of Red<br />

Valve Company, Carnegie, PA, U.S.A.)<br />

wiched between two rubber linings (or cartridges). They are often installed on the suction<br />

side of <strong>slurry</strong> pumps to provide a method of isolating them during repairs and maintenance.<br />

Most knife-gate valves are rated to a maximum of 1 MPa (150 psi), but some manufacturers<br />

offer valves rated at 2 MPa (300 psi).<br />

Globe valves are not suitable for <strong>slurry</strong> applications because they wear rather rapidly.<br />

To control <strong>slurry</strong> flows, a rubber pinch valve is recommended (Figure 2-23). The valve<br />

features a s<strong>pe</strong>cial reinforced sleeve. The sleeve is closed by pinching using a s<strong>pe</strong>cial roller<br />

(mechanical pressure) (Figure 2-24) or by the use of air pressure (Figure 2-25).<br />

Ceramic ball valves are used as shut-off valves for pi<strong>pe</strong>lines, particularly to close under<br />

high pressure.<br />

2-8 PRESSURE LOSSES THROUGH<br />

FITTINGS AT LOW REYNOLDS NUMBERS<br />

Certain <strong>slurry</strong> flows, particularly those of a non-Newtonian regime, do occur at relatively<br />

moderate Reynolds numbers and in laminar conditions (Tables 2-13 to 2-14). For many<br />

years, the method using the K factor and the equivalent length has been the most widely


FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

2.55<br />

FIGURE 2-25 Principles of o<strong>pe</strong>ration of a pneumatically actuated pinch valve. (Courtesy of<br />

Red Valve Company, Carnegie, PA, U.S.A.)<br />

accepted method. It is based on ex<strong>pe</strong>rimental data obtained usually in steel pi<strong>pe</strong>s at very<br />

high Reynolds numbers. As the Reynolds number is reduced closer to laminar flow, the K<br />

factor becomes inversely proportional to it. Since certain homogeneous slurries are sometimes<br />

pum<strong>pe</strong>d at relatively low Reynolds numbers, even quite close to the critical value, it<br />

is important to emphasize an alternative approach. Hoo<strong>pe</strong>r (1992) emphasized the limitations<br />

of this method and proposed a two-K method:<br />

K1 K� K = � + ��<br />

(2-40)<br />

Re 1 + 1/D1-in


2.56 CHAPTER TWO<br />

TABLE 2-13 Equivalent Length of Fittings for Friction Loss of Calculations for<br />

Single-Phase Turbulent Flow*<br />

Fitting Ty<strong>pe</strong> Length/Diameter Ratio<br />

Standard threaded elbow 90 degree 30<br />

Standard threaded elbow 45 degree 16<br />

Standard threaded elbow Long radius 90 degree—5 diameter<br />

bend as used in <strong>slurry</strong> plants<br />

16<br />

Mitre bend 15 degree bend 4<br />

30 degree bend 8<br />

45 degree bend 15<br />

60 degree bend 25<br />

75 degree bend 40<br />

90 degree bend 60<br />

Standard tee Through flow 20<br />

Through branch 60<br />

*Data from Ingersoll Rand (1977).<br />

where<br />

K1 = value of K at a Reynolds number of 1<br />

K� = value of K at high Reynolds numbers<br />

DI-in = internal pi<strong>pe</strong> diameter in inches.<br />

Values of these two constants are presented in Table 2-15.<br />

Regarding the equivalent length method, Hoo<strong>pe</strong>r (1992) wrote:<br />

TABLE 2-14 Dynamic Loss Factor K for Expansions and Contractions, where Loss =<br />

KV 2 /2g*<br />

Fitting Description Loss factor K<br />

Pi<strong>pe</strong> exit Projecting sharp edged, rounded 1.0<br />

Pi<strong>pe</strong> entrance Inward projecting 0.78<br />

Pi<strong>pe</strong> entrance (flush) Sharp edged 0.5<br />

Bellmouth fillet/diameter = 0.02 0.28<br />

Bellmouth fillet/diameter = 0.04 0.24<br />

Bellmouth fillet/diameter = 0.06 0.15<br />

Bellmouth fillet/diameter = 0.10 0.09<br />

Bellmouth fillet/diameter = 0.15<br />

and up<br />

0.04<br />

Reentry pi<strong>pe</strong> L/D = 65<br />

Sudden enlargements in pi<strong>pe</strong>s 2 2 K = (1 – d 1/d 2) Sudden contractions in pi<strong>pe</strong>s 2 2 K = 0.5(1 – d 1/d 2) Gradual enlargements in pi<strong>pe</strong>s � Less than 45 degrees 2 2 2<br />

K = 2.6 sin (�/2)(1 – d 1/d 2) � Larger than 45 degrees 2 2 2<br />

K = (1 – d 1/d 2) Gradual contractions in pi<strong>pe</strong>s � Less than 45 degrees 2 2 K = 0.8 sin �(1 – d 1/d 2) � Larger than 45 degrees 2 2 K = 0.5(1 – d 1/d 2)�(s�in� ��/2�)�<br />

*Data from Ingersoll Rand (1977).


FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

TABLE 2-15 Constants for the Two-K Method* (after Hoo<strong>pe</strong>r 1992)<br />

The equivalent-length method concept contains a booby trap for the unwary. Every<br />

equivalent length method has a s<strong>pe</strong>cific friction factor ( f ) associated with it, because<br />

the equivalent lengths were originally develo<strong>pe</strong>d from the K factor in the formula<br />

Le = KD/f. This is why the latest version of the equivalent length method (the<br />

1976 edition of the Crane Technical Pa<strong>pe</strong>r 410 ...pro<strong>pe</strong>rly requires the use of two<br />

friction factors. The first is the actual friction factor for the pi<strong>pe</strong> ( f ), and the second<br />

is a “standard” friction factor for the particular fitting ( f T). Thus the two-K method<br />

is as easy to use and as accurate as the updated equivalent-length method.<br />

The two-K method will be explored further in Chapter 5.<br />

2.57<br />

K1 at K� at very<br />

Fitting Description Ty<strong>pe</strong> Re = 1 high Re<br />

Elbows 90° Standard R/D = 1, screwed 800 0.40<br />

Standard R/D = 1, flanged/welded 800 0.25<br />

Long radius (R/D = 1.5), all ty<strong>pe</strong>s 800 0.20<br />

Mitered Elbow R/D = 1.5 1 weld 90° 1000 1.15<br />

2 welds 45° 800 0.35<br />

3 welds 30° 800 0.30<br />

4 welds 22.5° 800 0.27<br />

5 welds 18° 800 0.25<br />

45° Standard (R/D = 1.0), all ty<strong>pe</strong>s 500 0.20<br />

Long radius (R/D = 1.5), all ty<strong>pe</strong>s 500 0.15<br />

Mitered, 1 weld, 45° angle 500 0.25<br />

Mitered, 2 welds, 22.5° angle 500 0.15<br />

180° Standard R/D = 1, screwed 1000 0.60<br />

Standard R/D = 1, flanged/welded 1000 0.35<br />

Long radius (R/D = 1.5), all ty<strong>pe</strong>s 1000 0.30<br />

Tees Used as elbows Standard, screwed 500 0.70<br />

Long radius, screwed 800 0.40<br />

standard, flanged/welded 800 0.80<br />

Stub-in-ty<strong>pe</strong> branch 1000 1.00<br />

Run-through tee Screwed 200 0.10<br />

Flanged or welded 150 0.05<br />

Stub-in-ty<strong>pe</strong> branch 100 0.00<br />

Valves Gate, ball, plug Full line size, � = 1 300 0.10<br />

Reduced trim, � = 0.9 500 0.15<br />

Reduced trim, � = 0.85 1000 0.25<br />

Globe Standard 1500 4.00<br />

Globe Angle or Y-ty<strong>pe</strong> 1000 2.00<br />

Diaphragm Dam ty<strong>pe</strong> 1000 2.00<br />

Butterfly 800 0.25<br />

Check Lift 2000 10.0<br />

Swing 1500 1.50<br />

Tilting check 1000 0.50<br />

*Use R/D = 1.5 values for R/D = 5 pi<strong>pe</strong> bends, 45° to 180°. Use appropriate tee values for flowthrough<br />

crosses.


2.58 CHAPTER TWO<br />

2-9 THE BERNOULLI EQUATION<br />

The last few sections of this chapter examined the concept of friction and pressure losses.<br />

The presence of friction forces, changes in elevation between one point and another along<br />

the piping, the presence of a pump to add energy to the fluid, or a turbine to extract energy<br />

can all be expressed in terms of the extended Bernoulli’s equation:<br />

(E p + E v + E z) 1 + E A = (E p + E v + E z) 2 + E E + E f + E m<br />

2 U 1 P2 + � + Z2g + EA = � + � + Z2g + EE + Ef + Em 2<br />

� 2<br />

where subscripts 1 and 2 refer to points 1 and 2.<br />

Ep = P1/� = energy due to static pressure <strong>pe</strong>r unit mass<br />

2 U 1/2 = energy due to dynamic pressure <strong>pe</strong>r unit mass<br />

Z = location of point above a reference datum<br />

EA = energy added (e.g., by a pump) <strong>pe</strong>r unit mass<br />

EE = energy extracted (e.g., by a turbine) <strong>pe</strong>r unit mass<br />

Ef = Energy <strong>pe</strong>r unit mass due to friction losses<br />

Em = Energy lost due to fittings, <strong>pe</strong>r unit mass<br />

In USCS units.<br />

2-10 ENERGY AND HYDRAULIC GRADE<br />

LINES WITH FRICTION<br />

(2-41)<br />

When the total energy for flow in a pi<strong>pe</strong>line is plotted against distance, a profile called the<br />

energy gradient line is obtained. The energy drops with friction or extraction through a<br />

turbine, and increases by absorption from a pump.<br />

The hydraulic gradient is the sum of the pressure and the potential energies. The hydraulic<br />

gradient is therefore smaller than the energy gradient by the dynamic head (Figure<br />

2-26).<br />

2-11 FUNDAMENTAL HEAT TRANSFER<br />

IN PIPES<br />

In many areas of the world, mining is done in cold climates (Figure 2-27). Long tailing<br />

pi<strong>pe</strong>lines are exposed to wind, snow, and freezing conditions. In some oil–sand processes,<br />

tem<strong>pe</strong>rature is used to facilitate the pumping or separation of tar from sand. In other<br />

processes, hot slurries are fed to autoclave furnaces. The field of heat transfer is immense,<br />

but in the following paragraphs, some fundamentals will be reviewed. There are three<br />

main phenomena of heat transfer:<br />

1. Conduction<br />

2. Convection<br />

3. Radiation<br />

P 1<br />

� �<br />

U 2 2


EGL<br />

HGL<br />

Energy and Hydraulic Gradients<br />

For a pump<br />

FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

EA 2<br />

E v=<br />

V /2<br />

EGL<br />

2<br />

V 1 /2<br />

HGL<br />

Energy and Hydraulic Gradients<br />

For an expansion<br />

FIGURE 2-26 Energy and hydraulic gradients.<br />

2<br />

V 2 /2<br />

2.59<br />

FIGURE 2-27 The construction of mines may require pi<strong>pe</strong>lines that o<strong>pe</strong>rate in extremely<br />

cold environments. This water pi<strong>pe</strong>line was insulated and heat-traced for an Arctic environment.


2.60 CHAPTER TWO<br />

TABLE 2-16 Examples of Conductivity Range<br />

Range of conductivity K, Range of conductivity K,<br />

Material W/m °K Btu-ft/hr-ft 2 °F<br />

Insulators 0.03–0.21 0.02–0.12<br />

Nonmetallic Liquids 0.09–0.70 0.05–0.40<br />

Nonmetallic Solids 0.03–2.6 0.02–1.5<br />

Liquid metals 8.7–78 5.0–45<br />

Metallic alloys 14–120 8.0–70<br />

Pure metals 52–420 30–240<br />

2-11-1 Conduction<br />

Heat transfer by conduction occurs essentially by molecular vibration and movement of<br />

free electrons. As metals have more free electrons than nonmetals, they are better conductors<br />

of heat. Thermal conductivity, also known as thermal conductance, is a measure of<br />

the rate of heat transfer <strong>pe</strong>r unit thickness. Examples of conductivity range are presented<br />

in Table 2-16<br />

Thermal conductivity is a function of tem<strong>pe</strong>rature. For metals it decreases with tem<strong>pe</strong>rature,<br />

whereas for insulators it increases with tem<strong>pe</strong>rature. To simplify matters,<br />

it is common to assume the thermal conductivity at the average tem<strong>pe</strong>rature of 1 – 2(T 1 +<br />

T 2).<br />

2-11-2 Thermal Resistance<br />

Defining heat transfer power as Q t, thermal resistance is defined as<br />

R th = (2-42)<br />

where Qt is expressed in watts or Btu/hr.<br />

For a flat plate with a thickness path length L and an area A, and if heat transfer occurs<br />

by conduction and kth is the thermal conductivity of the material, the resistance factor Rth is:<br />

L<br />

Rth = (2-43)<br />

For a layer of insulation around a pi<strong>pe</strong>, this equation is expressed in terms of the inner<br />

and outer radius of the insulation layer:<br />

ln(RO/RI) Rth = � (2-44)<br />

2�kthL<br />

2-11-3 The R Value<br />

T1 – T2 �<br />

Qt<br />

� kthA<br />

One term commonly used by the industry is the thermal resistance <strong>pe</strong>r unit area or R value.


R Value = � RthA (2-45)<br />

Qt/A<br />

2-11-4 The S<strong>pe</strong>cific Heat or Heat Capacity C ��<br />

The s<strong>pe</strong>cific heat capacity is defined as the energy required to increase the tem<strong>pe</strong>rature of<br />

a unit mass by a unit degree and is calculated as<br />

Qt = C�m�T (2-46).<br />

2-11-5 Characteristic Length<br />

Characteristic length is defined as the ratio of the volume to its surface area and is calculated<br />

as<br />

V<br />

Lc = (2-47)<br />

2-11-6 Thermal Diffusivity<br />

Thermal diffusivity is a measure of the s<strong>pe</strong>ed of propagation of a s<strong>pe</strong>cific tem<strong>pe</strong>rature<br />

into a solid. The higher the diffusivity, the faster the material will reach a certain tem<strong>pe</strong>rature.<br />

Thermal diffusivity is calculated as<br />

where<br />

� e = thermal resistivity (�-cm or �-in)<br />

� = diffusivity (m 2 /s or ft 2 /hr)<br />

K th = conductivity (W/m-°K or Btu-ft/hr-ft 2 -°F)<br />

C � = s<strong>pe</strong>cific heat capacity (J/kg°K – Btu/lbm-°F)<br />

2-11-7 Heat Transfer<br />

FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

T 1 – T 2<br />

� As<br />

k th<br />

2.61<br />

� = � (2-48)<br />

�eC� Heat transfer is essentially a transmission of energy from one body to another in a <strong>pe</strong>riod<br />

of time. For this reason, it has the same unit as power in SI units, i.e., the watt. In USCS<br />

units Btu/hr is used. However, many equations ignore the time factor. Heat transfer <strong>pe</strong>r<br />

unit area qta is often used so that the total heat transfer Qt over an area A is calculated as<br />

Qt = qtaA Qt = mC��T (2-49)<br />

where<br />

m = the mass of the body<br />

�T = the tem<strong>pe</strong>rature change or power or rate of heat transfer<br />

The rate of heat transfer or power associated with the flow is expressed as<br />

Pwt = �QC��T


2.62 CHAPTER TWO<br />

Heat transfer can take different forms when <strong>slurry</strong> is stored in tanks, varies in thickness,<br />

or flows in pi<strong>pe</strong>s. In the northern climates, loss of heat can lead to frozen pi<strong>pe</strong>lines.<br />

In the hot climates, the heat absorbed from the sun leads to expansion of plastic lines and<br />

significant pi<strong>pe</strong> stresses.<br />

2-12 CONCLUSION<br />

In this chapter, some very important principles regarding water flows were introduced.<br />

Since water is the principal carrier of <strong>slurry</strong> mixtures, the tools develo<strong>pe</strong>d in this chapter<br />

such as hydraulic friction gradients and methods to correlate the friction velocity with the<br />

friction factor will be extensively used for pi<strong>pe</strong> flow and o<strong>pe</strong>n channel flow of heterogeneous<br />

mixtures (Chapters 4 and 6).<br />

This chapter discussed some s<strong>pe</strong>cific valve ty<strong>pe</strong>s such pinch, rubber sleeve, and check<br />

valves. These valves have their own ex<strong>pe</strong>rimental loss coefficients, which need to be obtained<br />

from manufacturers. This chapter presented the conventional K and the new two-K<br />

loss factors. The two-K factor as develo<strong>pe</strong>d by Hoo<strong>pe</strong>r is of particular importance for<br />

<strong>slurry</strong> flows at low Reynolds numbers. The engineer should therefore avoid the common<br />

pitfall of using published data on turbulent water flows for conventional waterworks<br />

valves when estimating the losses in a <strong>slurry</strong> system.<br />

2-13 NOMENCLATURE<br />

A Cross-sectional area of the flow<br />

As Surface area<br />

C Hazen–Williams roughness factor<br />

Cc Coefficient of contraction<br />

Cd Discharge coefficient<br />

C� S<strong>pe</strong>cific heat or heat capacity<br />

Cv Valve coefficient<br />

Cve Velocity coefficient<br />

din Pi<strong>pe</strong> diameter expressed in inches<br />

DH Hydraulic diameter = 4A/P<br />

Di Conduit inner diameter (m)<br />

Dij Inner diameter of the pi<strong>pe</strong> j<br />

E Energy <strong>pe</strong>r unit mass<br />

EA Energy added <strong>pe</strong>r unit mass<br />

EE Energy extracted <strong>pe</strong>r unit mass<br />

Ef Energy due to friction loss <strong>pe</strong>r unit mass<br />

Em Energy lost due to fittings <strong>pe</strong>r unit mass<br />

Ep Energy due to static pressure <strong>pe</strong>r unit mass<br />

Ev Energy due to dynamic pressure <strong>pe</strong>r unit mass<br />

Ez Potential energy <strong>pe</strong>r unit mass due to elevation above a reference point<br />

fD Darcy friction factor<br />

fN Fanning friction factor<br />

Fr Friction force<br />

F12 Force between points 1 and 2<br />

g Acceleration due to gravity (9.8 m/s2 )


FUNDAMENTALS OF WATER FLOWS IN PIPES<br />

2.63<br />

gc Conversion factor between slugs and lbm or 32.2 ft/sec2 h Spacing between plates<br />

Hf Head loss due to friction<br />

Hv Head loss in the Hazen–Williams formula<br />

kth Conductivity<br />

Kf Pressure loss of the fitting f<br />

L Length of conduit or pi<strong>pe</strong><br />

Lc Characteristic length<br />

Le Entrance length<br />

Lj Length of the conduit j<br />

m The mass of the body<br />

P Pressure<br />

Ppsi Pressure in psi<br />

Pwt Rate of heat power transfer<br />

Q Flow rate (m3 /s)<br />

Qgpm Flow rate expressed in US gallons <strong>pe</strong>r minute<br />

Qideal Ideal flow rate through an orifice as product of area and velocity<br />

qth Heat transfer <strong>pe</strong>r unit area<br />

r local radius<br />

RI Radius at the inner wall of the pi<strong>pe</strong>, or inner radius in an annular flow<br />

RH Hydraulic radius = area/<strong>pe</strong>rimeter<br />

R Resistance factor for thermal insulation<br />

Ri is the pi<strong>pe</strong> inner radius (at the inside wall of the pi<strong>pe</strong>)<br />

RO Outer radius in an anuular flow<br />

Rth thermal factor<br />

Re Reynolds number<br />

S Slo<strong>pe</strong> or head <strong>pe</strong>r unit length<br />

S.G. S<strong>pe</strong>cific gravity<br />

T Average tem<strong>pe</strong>rature<br />

TDH Total dynamic head that a pump is required to develop<br />

u Velocity of the flow at distance y<br />

U Average s<strong>pe</strong>ed of a flow outside the boundary layer<br />

Uf Friction velocity<br />

Umax Maximum s<strong>pe</strong>ed in the boundary layer<br />

VO Practical velocity across an orifice due to vena contracta<br />

Vth Theoretical velocity across an orifice<br />

W weight<br />

y + The relative distance from the wall in the boundary layer<br />

ZA Elevation of a point above a reference grade<br />

� Shear strain<br />

d�/dt Wall shear rate or rate of shear strain with res<strong>pe</strong>ct to time<br />

� Diffusivity<br />

� The thickness of the boundary layer �<br />

� Linear roughness (m)<br />

� Carrier liquid absolute or dynamic viscosity (usually expressed in Pascal-seconds<br />

or poise)<br />

� Pythagoras number (ratio of circumference of a circle to its diameter)<br />

� Duration of the shear for a time-de<strong>pe</strong>ndent fluid<br />

� Density in kg/m3 or slug/ft3 �e Thermal resistivity<br />

� Shear stress at a height y or at a radius r


2.64 CHAPTER TWO<br />

�w Wall shear stress<br />

� kinematic viscosity (defined as absolute viscosity divided by density)<br />

2-14 REFERENCES<br />

Hoo<strong>pe</strong>r W. B. 1992. Fittings, Number and Ty<strong>pe</strong>s. pp. 391–397 of The Piping Design Handbook,<br />

Edited by J. J. McKetta. New York: Marcel Dekker.<br />

Ingersoll Rand. 1977. The Cameron Hydraulic Handbook. Ner Jersey: The Ingersoll Rand Company.<br />

Johnson, M. 1982. Non-Newtonian Fluid System Design. Some Problems and Their Solutions. Pa<strong>pe</strong>r<br />

read at the 8th International Conference on the Hydraulic Transport of Solids in Pi<strong>pe</strong>, Johannesburg,<br />

South Africa.<br />

Lindeburg, M. R. 1997. Mechanical Engineering Reference Manual. Belmont, CA: Professional<br />

Publications Inc.<br />

Schlichting, H. 1968. Boundary Layer Theory, 6th ed. New York: McGraw-Hill.<br />

The Hydraulic Institute.1990. Engineering Data Book. Cleveland, OH: The Hydraulic Institute.<br />

Wasp E., J. Penny, and R. Handy. 1977. Solid–Liquid Flow Slurry Pi<strong>pe</strong>line Transportation. Aedermannsdorf,<br />

Switzerland: Trans Tech Publications.


Index terms Links<br />

A<br />

ABB wrap-around motors 1.25<br />

Abrasion, (see also white cast iron) 8.3<br />

ACME thread 8.49 8.50<br />

Activators 7.39<br />

Adsorption 7.38<br />

Affinity laws 8.60<br />

Agitators (see also mixers)<br />

anchor 7.43 7.44<br />

biological treatment 7.44<br />

critical s<strong>pe</strong>ed 7.45<br />

forces 7.45 7.47 7.54 7.55 7.56 7.57<br />

7.59<br />

helicoidal 7.44<br />

pump 8.56<br />

pro<strong>pe</strong>ller-ty<strong>pe</strong> 7.43 7.47<br />

Air release valve 11.22 11.23<br />

Albertson sha<strong>pe</strong> factor 3.12 3.13 3.14 3.15 3.16 3.17<br />

Alkination 11.10<br />

Allen’s equation 3.6<br />

Alundum 10.21<br />

Anderson 8.6<br />

Andesite 7.22<br />

Angularity test 1.11<br />

Antemina 11.31<br />

Anti-dunes 6.30 6.32 6.33<br />

Apatite 7.39 11.17<br />

API (American Petroleum Institute)<br />

Grade 5LX65 11.29<br />

Grade 5LX70 11.29<br />

RP1102 11.28<br />

RP1109 11.28<br />

RP1111 11.28<br />

Std 1104 11.29<br />

Std 1130 11.29<br />

Archimedean number 4.11 4.13 4.51<br />

Area ratio 8.7<br />

Argentina 1.3 1.34 11.22<br />

ASARCO 1.25<br />

Ash 10.21<br />

ASTM Standards<br />

A532 10.6 10.10<br />

B3 1.11 11.28


Index terms Links<br />

D422 12.21<br />

D698 12.22<br />

D854 12.21<br />

D1556 12.21<br />

D2488 1.13<br />

D3017 12.21<br />

D3155 1.8<br />

D4767 12.22<br />

Aswan High Dam 1.35<br />

Athabasca region 11.23<br />

Austenitic 10.9<br />

Australia 1.34 1.35 11.10 11.11<br />

Australian standard AS2576 11.19<br />

B<br />

Backfill 11.24 11.25 11.26<br />

Bajo Alumbrera 1.3 11.22<br />

Ball mill (see mill)<br />

Barite 7.17 7.22<br />

Basalt 7.22<br />

Batu Hijau 1.34<br />

Bauxite 3.19 5.3 5.4 5.7 5.35 5.36<br />

5.38 7.17 7.22 10.21 11.1<br />

Bearings, pump 8.2 8.3<br />

Bed layer 6.18<br />

Bedforms (see dunes)<br />

Belonovo.Novosibirsk (Pi<strong>pe</strong>line) 1.34 11.10<br />

Bentonite 11.22<br />

Bernoulli’s equation 2.58<br />

BHRA or BHR Group 4.2<br />

Bingham plastics 3.17 3.18 3.19 3.20 3.21 3.22<br />

5.2<br />

5.11 6.4<br />

Bitumen (see also oil sands, tar) 11.24<br />

Black Mesa Pi<strong>pe</strong>line 1.34 11.2 11.7<br />

Blasius’s equation 2.35<br />

Blockage<br />

Bolt adjustement for pump bearing<br />

4.43 4.45 4.46<br />

cartridge 8.3<br />

Bond equation 7.1<br />

Bougainville Pi<strong>pe</strong>line 1.34<br />

Boundary Layer 2.45<br />

Brazil 1.34<br />

Brazilian split test<br />

British Standard<br />

1.11<br />

BS 812 1.11


Index terms Links<br />

BS 5930 1.8<br />

Buckingham equation 3.34 5.5 5.6<br />

Budryck’s equation 3.7 3.8 3.9<br />

Buffer layer 2.45<br />

Buoyancy 3.3<br />

C<br />

Calavaras Pi<strong>pe</strong>line 1.34<br />

Calcium carbonate 14<br />

California 1.3<br />

Cameron 2.50<br />

Canada 11.23 11.24<br />

Canal <strong>pe</strong>rmissible velocity 6.25<br />

Carbides 10.10<br />

Carbon 10.21<br />

Carbonate 1.8<br />

Carborundum 10.21<br />

Carman, Kozeney equation 3.10<br />

Carson slurries 3.29 3.37<br />

Cascades 6.54 6.55 6.56<br />

Cast iron 10.3 10.4 11.10<br />

Cavitation<br />

Celik and Rodi, method for dilute<br />

8.18 8.20 8.22<br />

8.24 8.25<br />

flows in launders 6.18 6.21<br />

Cellulose acetate 3.21<br />

Cement, kiln feed <strong>slurry</strong> 1.16 3.19 5.3 5.4 5.7<br />

Cement clinker 7.22<br />

Centrifuges 7.12 7.62 7.63 7.64<br />

Chalcopyrite 7.38<br />

Chemineer scale, for agitation 7.49 7.50 7.51<br />

Chemisorption 7.39<br />

Chevron Phosphate Pi<strong>pe</strong>line 1.34 11.19 11.20<br />

Chezy number 6.4 6.9 6.32<br />

Chile<br />

Chilton and Stainsby method for<br />

1.3 11.22<br />

non-Newtonian flows 5.19 5.20 5.21 5.22<br />

China 1.33 1.34<br />

Chokes 2.12 11.20 12.22<br />

Chongin Pi<strong>pe</strong>line 1.34<br />

Churchill’s equation 4.52 5.30<br />

Clarification 1.5 1.31<br />

Classification<br />

Classifier<br />

1.25 1.26<br />

hydraulic<br />

hydrocyclone (see<br />

hydrocyclone)<br />

7.32


Index terms Links<br />

mechanical 1.26 7.33<br />

Clay 1.12 1.15 1.16 1.18 1.34 3.19<br />

3.32 5.4<br />

5.7 5.31 5.38 5.39<br />

Clift sha<strong>pe</strong> factor 3.13 3.14 3.15<br />

Closure and reclamation plan 12.23 12.24<br />

Coal 1.17 1.32 1.34 3.10 3.19 3.32<br />

4.26 7.17 10.21 11.1<br />

degradation 11.3 11.4<br />

dewatering 11.6 11.7<br />

ship loading 11.8 11.9<br />

size of particles for pumping 11.2 11.3<br />

tailings 3.19<br />

terminal velocity 3.10<br />

Coal-magnetite mixture 11.4 11.5<br />

Coal-oil mixture<br />

11.5 11.6<br />

Coal-water mixture, combustion 11.8 11.9 11.10<br />

Coalescence 1.19<br />

Cobbles 6.25<br />

Codelco 11.22<br />

Coefficient of curvature 1.13<br />

Coefficient of uniformity 1.13<br />

Coke 7.22<br />

Colebrook’s equation 2.9 2.10<br />

Collahausi Pi<strong>pe</strong>line 1.34<br />

Colorado School of Mines 1.32<br />

Compound mixtures 4.33<br />

Concentrate<br />

Concentration<br />

1.27 1.30<br />

volume 1.19 1.20 1.21 1.22<br />

weight 1.19 1.21<br />

Conduction 2.60<br />

Conductivity 2.60<br />

Conduits, pressure losses 2.44 2.46 2.47<br />

Consolidated Coal Company 1.32 4.30<br />

Consolidation Pi<strong>pe</strong>line 1.34<br />

Contact load<br />

Cop<strong>pe</strong>r<br />

4.51<br />

concentrate <strong>slurry</strong> 1.16 1.32 1.34 3.19 10.21 11.21<br />

ore 1.34 7.22<br />

tailings 1.34<br />

Cost estimates 12.24 12.27<br />

Coulombic friction 4.51<br />

Critical depth, in o<strong>pe</strong>n channels 6.59<br />

Critical slo<strong>pe</strong> 4.59<br />

Critical s<strong>pe</strong>ed 11.14


Index terms Links<br />

Critical velocity, in o<strong>pe</strong>n channels 6.23<br />

Crushing 1.3 1.24 1.25 7.3 7.11<br />

Crushers<br />

cone<br />

gyratory 7.7 7.8<br />

impact 7.8 7.10<br />

jaw 7.5 7.6 7.7<br />

primary 1.25 7.4<br />

roll 7.11<br />

secondary 1.25 7.9<br />

tertiary 1.25<br />

Cuajone 1.30<br />

Cunningham Creek Dam 1.35<br />

Cutwater, pump 8.29<br />

Cyclone (see also hydrocyclones) 1.25 1.28 1.29<br />

D<br />

Dallas White Rock Pi<strong>pe</strong>line 1.34<br />

Darcy, Weisbach equation 2.45<br />

Darby<br />

method for Bingham plastics 5.9 5.10<br />

method for yield-pseudoplastics 5.24 5.25 5.26 5.27 5.28<br />

Density 1.8 1.19 1.20<br />

Deposited bed 4.56<br />

Dewatering 1.30 11.6 11.7 11.10<br />

Diffuser 8.13<br />

Diffusivity, thermal 2.48<br />

Dilatancy 1.12<br />

Dilatant slurries 3.17 3.28<br />

Diorite 7.22<br />

Discharge, coefficient of 2.49 2.50<br />

Dithiosphophate 7.39<br />

Dodge and Metzner model 5.22<br />

Dolomite 7.17<br />

Dominguez’s equation 6.26 6.27<br />

Drag<br />

coefficient 3.2 3.3 3.4 3.5 5.39 5.40<br />

force 3.1 3.2 7.45<br />

reduction in non-Newtonian<br />

flows 5.39<br />

Dredge, arm 1.7<br />

Dredgeability 1.15<br />

Dredging 1.5 1.7 1.34 1.36 8.6 12.6<br />

sand cutter 1.5 1.9<br />

Drillability, test 1.11<br />

Drive end, pump 8.2 8.42 8.52<br />

Drop boxes 6.55 6.56 6.71


Index terms Links<br />

Duboy’s equation 6.11 6.36<br />

Ductile iron 10.4<br />

Dunes 6.30 6.32 6.33<br />

Durand and Condolios’s equations 1.32 4.1 4.20<br />

velocity factor 4.8<br />

Dust, blast furnace 10.21<br />

Dynamic seal (see ex<strong>pe</strong>ller)<br />

E<br />

Efficiency, hydraulic 8.5 8.18 8.22<br />

Efficiency ratio, pump 8.64 8.72<br />

Egypt 1.3 1.6 1.32 1.34<br />

Einstein’s equation 1.21 6.18 6.36<br />

Einstein, Brown’s equation 6.36<br />

Elasticity, static modulus 1.11<br />

Electrometallurgy 1.24<br />

Emergency pond 12.9<br />

Emulsions, 1.16 1.17 1.19<br />

Entrance length 2.45<br />

Equivalent length of fittings 2.45 2.56<br />

Erosion, corrosion 12.19<br />

Escondida 1.3 1.34 11.22<br />

Eskay Creek 11.2<br />

Ethiopia 1.34<br />

Ethyl xanthate 7.38<br />

ETSI Pi<strong>pe</strong>line 1.34 11.2<br />

Eutectic 10.10<br />

Ex<strong>pe</strong>ller 8.2 8.5 8.34<br />

F<br />

Fanning friction (see friction factor)<br />

Feldspar 7.17 7.22<br />

Ferrite 10.10<br />

Ferrochrome 7.22 10.7<br />

Ferro-manganese 7.22<br />

Ferro-molybdenum 10.6<br />

Ferro-silicon 7.22<br />

Filtering 1.27 1.30<br />

Fittings 2.54 2.55 2.56 2.57<br />

5.3 5.4 5.75<br />

Flange loads 8.52<br />

Flint 7.17<br />

Flocculation launders 6.44<br />

Flocculants 3.24 7.38 12.5<br />

Flotation 1.3 1.25 1.26 1.27 1.28<br />

1.29 7.12<br />

froth 1.28<br />

7.38


Index terms Links<br />

Flows<br />

heterogeneous 1.16<br />

homogeneous 1.16<br />

laminar 1.19<br />

transition 2.3<br />

turbulent 2.3<br />

Flue gas desulphurisation (FGD) 10.9<br />

Flyash 3.21 10.21<br />

Fluorspar 7.17 7.22<br />

Fracture toughness 10.10<br />

Francis’s equation 3.8<br />

Freeport 1.34<br />

Friction<br />

factor for single-phase fluids 1.19 2.4 2.12<br />

factor for heterogeneous flows 4.15<br />

force 2.2<br />

gradient 2.24 2.44<br />

Of heterogeneous flows 4.19 4.41 6.29 6.39<br />

velocity 2.35 4.30 6.30 6.34<br />

Frost and Nielsen’s equation 8.22<br />

Froth (see also flotation) 3.18 7.39 8.2 8.56 8.57 8.68<br />

8.70<br />

8.72<br />

Frothers 7.39<br />

Froude number 4.11 6.45 6.67<br />

G<br />

G-gradient 6.44<br />

Gabbro 7.17 7.22<br />

Gangue 1.24 1.26<br />

Garbett and Yiu 11.10<br />

Geho pumps 9.8 9.13 11.22<br />

Geller and Gray 4.25<br />

Gilles equation 4.11 4.13 4.15<br />

Gioasfertil Phosphate Pi<strong>pe</strong>line 11.20<br />

Gladstone Pi<strong>pe</strong>line 11.10 11.13<br />

Gland, pump 8.2<br />

Glass 7.17 7.22<br />

Gneiss 7.22<br />

Gold<br />

ore 7.22<br />

tailings 3.19 11.2<br />

Gradient<br />

energy (see friction gradient) 2.59<br />

hydraulic 2.59<br />

Graf and Acaroglu’s equation 4.45 4.46 6.33 6.34<br />

Granite 7.17 7.22


Index terms Links<br />

Graphite 1.28 3.21 7.22<br />

Gravel 1.15 3.10 6.25 7.22<br />

Grease cup, pump 8.2<br />

Green 6.45<br />

Grey iron 10.3<br />

Grinability work index 7.14 7.17<br />

Grinding 1.24 1.25 1.26 1.30 7.12 7.31<br />

Gypsum rock 7.17 7.22 10.21<br />

H<br />

Hagen, Poiseville equation 2.6<br />

Hanks and Dadia equation 5.9<br />

Hanks and Pratt equation 5.8<br />

Hanks and Ricks, method for yield<br />

pseudoplastics 5.17 5.18<br />

Hardness of particles 1.3 1.10 10.3 11.17 A.1<br />

Hatch and Associates 1.30<br />

Hayden and Stelson equation 4.29<br />

Hazen, William’s method 2.18 2.19 2.20<br />

Head coefficient 8.14<br />

Head loss, in o<strong>pe</strong>n channels 6.3<br />

Head, pump 8.5 8.8 8.10 8.25 8.26<br />

Head ratio, pump 8.64 8.72<br />

Heat<br />

capacity 2.61<br />

transfer 2.61<br />

Hedstrom number 5.1 5.2 5.6<br />

Hematite (see also iron ore) 7.22 7.39<br />

Herschel, Bulkley model 5.19 5.21<br />

viscosity 5.19<br />

Heterogeneous flows 1.16<br />

Heywood’s equation 5.14 5.18<br />

High chrome alloy (see also 8.4 10.5 10.10<br />

white cast iron)<br />

High-density polyethylene (HDPE) 1.31 2.19 2.20 11.8 11.22<br />

Hindustan Zinc Phosphate Pi<strong>pe</strong>line 11.31<br />

Homogeneous flows 1.16<br />

Hoses 10.15<br />

Hunt’s equation 4.31<br />

Hydraulic diameter 6.3<br />

Hydraulic friction gradient of<br />

heterogeneous flow 4.9 4.19 4.25<br />

Hydraulic friction gradient of water 2.19 2.20 2.22 2.44 4.9<br />

Hydraulic grade line 2.58 2.59<br />

Hydraulic jump 6.60 6.63<br />

Hydraulic radius 6.2 6.3


Index terms Links<br />

Hydrocyclone 1.25 1.28 1.29 7.12 7.23 7.33<br />

7.38 11.7 11.25<br />

Hydrolic degradation 10.17<br />

Hydrometallurgy 1.22<br />

Hy<strong>pe</strong>rchrome 11.19<br />

Hy<strong>pe</strong>reutectic<br />

I<br />

10.10<br />

Impactors 11.22<br />

Im<strong>pe</strong>ller 8.2 8.7 8.18 8.31 8.34 8.42<br />

Inclined flows 4.58 4.61<br />

India<br />

Inter-Agency Committee on Water<br />

11.21<br />

Resources 3.14<br />

International Society for Rock<br />

Mechanics<br />

1.11<br />

Inverted U-flowmeter<br />

Iron (see also cast iron)<br />

4.57<br />

concentrate 1.28 1.32 1.34 10.21<br />

ore 7.22 10.21 11.12<br />

ore, hematite 7.22<br />

ore, oolitic 7.22<br />

ore, s<strong>pe</strong>cular 7.22<br />

ore, taconite 7.22<br />

oxide 5.3 5.4 5.7<br />

sand 11.1<br />

Irvine, method for pseudoplastics 5.16<br />

Ismail equation for sediment<br />

distribution<br />

J<br />

4 30<br />

Jian Shan Phosphate Pi<strong>pe</strong>line<br />

K<br />

K-factor<br />

1.34<br />

for conduits in turbulent flows 2.45 2.46 2.47 2.56<br />

at low Reynolds Numbers 2.54<br />

5.74<br />

2.55 2.56 2.57 5.3 5.4<br />

Kaolin (see also clay) 3.19 3.21 3.22 3.23 5.38 5.39<br />

9.7 10.21<br />

Karasev function 6.41<br />

Kenya 1.4<br />

Kimbelite tails 3.19<br />

Khan and Richardson model for<br />

two<br />

layers 4.53<br />

Knelson, concentrators 7.63<br />

Koorawatha dam 1.35


Index terms Links<br />

Korrumbyn Creek Dam 1.35<br />

Kozney’s equation 3.11 3.12<br />

KSB-GIW, <strong>slurry</strong> research lab 1.19<br />

L<br />

Laminar flows, single phase 2.6 2.8<br />

La Parla, Hercules Pi<strong>pe</strong>line 1.34 11.15<br />

Larox valves 11.22<br />

Las Truchas 1.34<br />

Launders (see also o<strong>pe</strong>n 10.14 10.15<br />

channel flows)<br />

Leaching<br />

acid 11.27<br />

alkaline 11.27<br />

Lead, ore 7.22<br />

Lead, zinc, ore 7.22<br />

Length, equivalent 2.56 2.57<br />

Lift<br />

force 3.1 3.2 3.10 4.25 7.45<br />

coefficient 3.2<br />

Limestone <strong>slurry</strong> 1.16 1.34 3.19 5.3 5.7 7.17<br />

7.22 10.21 11.1 11.1 1.13<br />

Liminite 3.19 5.3 5.4 5.7<br />

Limonite 10.21<br />

Liners<br />

clay 11.27<br />

geological 11.27<br />

mantle 10.10<br />

pump (see also rubber, white 8.2 8.27 10.10 10.17<br />

cast iron)<br />

synthetic 11.27<br />

Los Bronces 1.34<br />

M<br />

Magnesite 7.22<br />

Magnesium hydroxide 3.21<br />

Magnetite 7.17 10.21<br />

Malleable iron 10.4<br />

Manganese ore 7.22<br />

Manning number 6.4 6.5 6.6 6.7<br />

Martensitic 10.9<br />

Materials selection 10.1 10.21 10.19<br />

Mazdak pump 8.3 8.30 8.54 8.56 8.58 8.59<br />

Mazdak, Slurry Research Lab 1.19<br />

Melbourne University, Australia 1.19<br />

Metallurgy, extractive 1.22<br />

Metals, stress, strain 10.1 10.3


Index terms Links<br />

Metso Minerals (formerly the 7.4 7.13<br />

companies of Nordberg and<br />

Svedala)<br />

Metzner 1.22 1.24<br />

Metzner, Reed, method for<br />

pseudoplastics 5.11 5.13<br />

Meyer Peter curve 6.36<br />

Meyer Peter Muler curve 6.36<br />

Michigan Limestone Tailings<br />

pi<strong>pe</strong>line 1.34<br />

Mill<br />

autogeneous 1.25 1.30<br />

ball 1.25 1.30 7.12 7.21 7.23 7.26<br />

critical s<strong>pe</strong>ed 7.25<br />

grate 7.24<br />

hammer 7.24 7.28 7.31<br />

liners 7.21 7.23 7.25<br />

overflow 7.24<br />

<strong>pe</strong>bble 1.25 7.23 7.24<br />

rod 1.25 7.19 7.23 7.26<br />

roller 7.28<br />

semi. Autogenous (SAG) 1.25 1.26 1.29 7.23 7.27<br />

spindle 7.24 7.28<br />

tower 7.24 7.28<br />

trommels 7.25<br />

trunnions 7.23<br />

tumbling 7.23<br />

vertical 1.30 7.24 7.28<br />

vibrating 7.24 7.28<br />

Miller number 9.1 10.20 10.22 11.18<br />

Milling (see also mills) 1.3 1.24 1.24<br />

Mixers 7.4 7.59<br />

correction factor 7.49 7.50<br />

diameter sizing 7.51<br />

equivalent volume 7.48<br />

flow coefficient 7.48 7.49<br />

power coefficient 7.48 7.49<br />

Reynolds number 7.48<br />

Mixing length 6.9<br />

Moballoy, white iron 10.10<br />

Mobile Pulley and Machine Works 1.7 1.8 1.9 10.10<br />

Mogas valves 11.22<br />

Molerus diagram 5.26<br />

Molybdenum 1.28 7.22<br />

Montuori number 6.55 6.56 6.57


Index terms Links<br />

Moody diagram 2.10 2.11 5.29<br />

Motor, wrap.around 1.25<br />

Mud (see also red mud)<br />

drilling 1.16 3.21 10.21 11.22<br />

Munroe’s equation 3.8<br />

N<br />

Natural <strong>slurry</strong> flows 1.4 1.5<br />

Newtonian flow 3.17<br />

New Zealand Sands 1.34<br />

Newitt 4.1 4.2 4.28 4.29 4.44<br />

Nickel 4.28 7.22<br />

Ni-hard alloys (see also white 8.4 10.5 10.8 11.10<br />

cast iron)<br />

Nikrudase roughness 5.3 5.4 5.71 6.34<br />

Nile River 1.3 1.5<br />

Non-Newtonian flows 1.16 2.49 3.17 3.38 5.1 5.42<br />

7.54 11.12<br />

Nordberg (see Metso Minerals)<br />

Nordstrom valves 11.22<br />

Nozzle 2.49<br />

NPSH 8.18 8.20 8.22 8.23 8.24 8.25<br />

O<br />

Ohio Pi<strong>pe</strong>line 11.7<br />

Oil sands, (see also tar sands) 11.4 11.23 11.24<br />

Oil shale 7.22<br />

OK Tedi 1.34 11.21<br />

O<strong>pe</strong>n channel flows 1.29<br />

Ore, classification 1.24 1.26<br />

Orifice plate 2.49 2.50<br />

Orimulsion 1.17 5.29<br />

P<br />

Particle sizes conversion 1.14<br />

Peat 1.15<br />

Pechuka tanks 7.42<br />

PDVSA, Bitor 1.17<br />

Pena pi<strong>pe</strong>line 1.34<br />

Permanent International<br />

Association 1.5 1.10 1.11 1.12<br />

of Navigation Congresses<br />

Pi<strong>pe</strong><br />

concrete 2.10 2.12<br />

high-density polyethylene 2.10 2.19 2.20<br />

rubber-lined 2.10 2.12 2.16 2.17 2.18<br />

steel 2.10 2.13<br />

2.15<br />

Pi<strong>pe</strong>lines<br />

Pinto Valley<br />

1.27<br />

1.34<br />

1.29 1.31 1.34 11.1 11.31


Index terms Links<br />

Phosphate<br />

maton 11.18<br />

matrix 10.9 10.21 11.15<br />

ore concentrate 1.34 11.1 11.15<br />

rock 7.22 12.6<br />

tailings 3.19<br />

Phosphoric acid 10.9 11.15 11.19<br />

Placer mining 1.15<br />

Plasticity index 1.15<br />

Point load test 1.11<br />

Poiseville flow 2.3 2.6<br />

Porosity 1.8<br />

Potash ore 7.22<br />

Prandtl, Colebrook’s equation 2.10<br />

Pressure<br />

dynamic 2.4<br />

gradient for heterogeneous<br />

flows 4.19 4.41<br />

loss due to friction 2.45<br />

loss in fittings at low Reynolds<br />

numbers 5.3 5.4 5.74 5.75<br />

pump 8.28 8.29<br />

start-up 5.2 5.5<br />

Pseudohomogeneous flows 4.18 4.19 4.47 4.48<br />

Pseudoplastics 3.17 5.11 5.17<br />

yield 3.17 5.17 5.22 5.24 5.28<br />

Pulley, radial force 8.51 8.52<br />

Pulp and pa<strong>pe</strong>r, friction calculations 5.40 5.41<br />

Pumps 1.28 1.36 8.1 8.72 9.1 9.16<br />

chop<strong>pe</strong>r 8.59<br />

diaphragm 9.8 9.9 9.10 9.11 9.12 9.13<br />

dredge 8.2 8.59 8.60 10.10<br />

froth-handling 8.2<br />

lockhop<strong>pe</strong>r 9.15 9.16<br />

low-head 8.2<br />

mill discharge 8.2<br />

<strong>pe</strong>rformance corrections 8.61 8.62 8.72<br />

<strong>pe</strong>ristaltic 9.13<br />

piston 9.1 9.2 9.3 9.4 9.5<br />

plunger 9.6 9.7 9.8 11.15 11.20 11.21<br />

positive displacement 1.31 9.1 9.16 11.24<br />

rotary lobe 9.14 9.15<br />

submersible 8.2 12.9 12.10<br />

tailings 8.2<br />

tank 8.57


Index terms Links<br />

unlined 8.3<br />

vertical cantilever 8.2 8.53 8.59<br />

Pyrhotite, ore 7.22<br />

Pyrite 7.22 10.21<br />

Pyrometallurgy 1.24<br />

Q<br />

Quartz 7.17 7.22<br />

Quartzite 7.17 7.22<br />

Quipolly Reservoir 1.35<br />

R<br />

R-value 2.60 2.61<br />

Rabinowitsch, Mooney, 5.11<br />

method for pseudoplastics<br />

Radial, force 7.47<br />

Radium 11.27<br />

Reagents 1.28<br />

Recirculation, in im<strong>pe</strong>llers 8.12<br />

Recirculation load 7.21<br />

Reclaim water 12.6<br />

Red mud 3.19 5.35 5.37<br />

Re<strong>pe</strong>ller (see ex<strong>pe</strong>ller)<br />

Resistance, thermal 2.60<br />

Reynolds Number<br />

single phase fluids 2.3 2.5<br />

critical 2.4 5.8 5.14 5.15 5.18<br />

modified 5.11 5.14 5.20 5.25<br />

particle<br />

Rheology 3.32 3.38 12.5 12.6<br />

Rheo<strong>pe</strong>ctic slurries 3.17 3.30<br />

Richards’s equation 3.6<br />

Rittinger’s equation 3.7 3.9<br />

Rod mills 7.19<br />

Roughness<br />

absolute 2.10 2.11 2.12 6.6 6.7 6.31<br />

effects on non-Newtonian<br />

flows 5.29 5.3 5.4 5.73<br />

equivalent 6.25<br />

piping material 2.10 2.11 2.12 6.31<br />

relative 2.13 2.18<br />

sand 4.52<br />

Rubber<br />

armadillo 10.14<br />

Buna-N 2.53<br />

carbon-black-filled 10.13<br />

carbon-black and silicon-filled 10.13<br />

carboxylic nitrile 10.17


Index terms Links<br />

chlorobutyl 2.53<br />

EPDM 2.53 10.15<br />

food grade 10.12 10.15<br />

fluoro-elastomer 10.15<br />

Hypalon 2.53 10.15 10.17<br />

jade green armabond 10.14<br />

lining 7.19 7.20 7.35 8.4<br />

natural 10.11 10.13<br />

natural aashto 10.11<br />

Neoprene 2.53 10.14<br />

nitrile 10.15<br />

polychlorene 10.13<br />

polyurethane 10.16 10.18 11.26<br />

pure tan gum 2.53 10.11<br />

synthetic 10.13 10.17<br />

viton 2.53 10.18<br />

white food grade 10.12<br />

Rubber/butadiene styrene, 10.12<br />

graphite-filled<br />

Rugby Pi<strong>pe</strong>line 1.34<br />

Russia 1.34 11.10<br />

Rutile, ore<br />

Ryan and Johnson, method for<br />

7.22<br />

pseudoplastics<br />

S<br />

SAG (see semiautogeneous mill)<br />

5.14 5.15<br />

Saltation 4.3 4.43 4.47<br />

Samarco 1.34 11.12<br />

Sand 1.15 1.18 6.25 10.21<br />

mineral 1.28<br />

Sandvik 7.7 7.8<br />

Sandstone 7.17<br />

Saskatchewan Science Research<br />

Center<br />

1.19 4.53 11.24<br />

Savage River 1.34<br />

SCADA 11.29 12.1 12.21<br />

Schiller’s equation<br />

Screens<br />

4.9 4.11 4.46<br />

banana 7.32<br />

shaking 7.32<br />

trommel 7.32<br />

vibrating 7.32<br />

Screening devices 7.31<br />

Sedimentation 1.5 1.34 7.59<br />

gravity 7.60 7.62


Index terms Links<br />

Seismic velocity test 1.11<br />

Semiautogeneous mill 7.12 7.18<br />

Separation<br />

electrostatic 1.2 1.26 1.27 1.28 1.29<br />

gravity 1.25 1.26 1.27<br />

magnetic 1.25 1.26 1.27 1.28 7.12 7.38<br />

Settling design s<strong>pe</strong>ed for mixers 7.49<br />

Sewage 3.11 5.3 5.4 5.7 7.53 10.21<br />

Shale 6.25 7.17 7.23<br />

Shear rate 1.15 2.7 3.18 3.38 5.2 5.9<br />

5.14 5.19 5.3 5.4 5.75 7.53<br />

Shear stress of flows 2.2 2.8<br />

Shook’s bimodal model 4.50<br />

Shut-down of pi<strong>pe</strong>line 4.45<br />

Siemens wrap-around motors 1.25<br />

Sierra Grande 1.34<br />

Sieve diameter 3.16<br />

Silica 7.22 7.39<br />

sand 7.23 10.21<br />

Silicon carbide 7.23<br />

Silt 1.4 1.5 1.12 1.13 1.15 1.18<br />

1.36 6.25<br />

Siltation 1.3 1.34 1.35 1.36 1.37<br />

Simplot 1.34<br />

Slack flow (see o<strong>pe</strong>n channel flows)<br />

Slag 7.22<br />

Sleeve, shaft 8.2<br />

Slip of coarse materials 6.35<br />

Slippage, wall 5.3 5.4 5.73<br />

Slip factor, in pumps 8.11 8.13<br />

Slip of coarse materials 6.35 8.61 8.64<br />

Slug flow 6.55 6.56<br />

Sodium silicate 7.23<br />

SOGREAH 1.32 4.2<br />

Soils<br />

classification 1.5 1.6 1.11<br />

coefficient of curvature 1.13<br />

coefficient of uniformity 1. 13<br />

cohesive 1.5 1.12<br />

composition tests 1.8<br />

liquid limit 1.13<br />

non.cohesive 1.5<br />

organic 1.12<br />

particle sizes 1.13<br />

plasticity 1.13 1.15


Index terms Links<br />

stratification 1.10<br />

strength 1.10 1.12<br />

testing 1.8 1.12 1.15<br />

textures 1.13<br />

Southern Peru Cop<strong>pe</strong>r 1.34<br />

S<strong>pe</strong>cific heat 1.22<br />

S<strong>pe</strong>cific s<strong>pe</strong>ed, pump 8.14 8.18<br />

Spodumene, ore<br />

Stainless steel 10.9<br />

Stepanoff 8.6<br />

Steward 11.24 11.25<br />

Stockpile 1.24 7.11<br />

Stoke’s equation 3.5 3.6<br />

Stratification velocity 4.49<br />

Stratified flows 4.48 4.50<br />

Stuffing box, pump 8.2<br />

Suction mouthpiece, for dredging<br />

boats 1.8<br />

Suction plate, pump 8.2<br />

Sudan 1.4<br />

Suez Canal 1.32<br />

Sulfur 10.21<br />

Sulzer 1.18<br />

Svedala (see Metso Minerals)<br />

Swamee, Jain friction factor 2.9<br />

Swirl number 7.33<br />

Syenite 7.23<br />

T<br />

Taconite (see also iron ore) 1.16 1.24 4.14<br />

7.17 7.22 11.15<br />

Tailings 1.3 1.10 1.25 1.27 1.28 1.30<br />

8.7 10.21 12.1 12.3 12.28<br />

dam 12.11 12.18 12.21<br />

submerged disposal 12.15 12.17<br />

Tar (see also bitumen) 5.29<br />

Texas A&M University 1.19<br />

Thermal conductivity 1.22<br />

Thickeners 7.60 7.61 7.62 12.2 12.5 12.6<br />

Thickening 1.26 1.27 1.30<br />

Thixotropic slurries 3.17 3.30 3.32 5.28 5.29<br />

Thomas’s equation 1.21 1.22<br />

Thorium oxide 3.23<br />

Thread pull force 8.48 8.51<br />

Throatbush, pump<br />

Thrust<br />

8.29 8.30<br />

axial 8.42 8.48


Index terms Links<br />

radial 8.42 8.48<br />

Tin, ore 7.23<br />

Titanium<br />

ore 7.23<br />

dioxide 3.23<br />

Tomb chart, pumps 8.6<br />

Tomita, method for pseudoplastics 5.13 5.16<br />

Torrance, method for yield 5.18 5.19 5.29<br />

pseudoplastics<br />

Trap rock 7.23<br />

Transition flows (single phase 2.8 2.9<br />

liquids)<br />

Turbine, <strong>slurry</strong> 1.31<br />

Turbulent layer 2.45<br />

Turton and Levenspiel equation 3.4<br />

Two-layer model 4.50 4.56<br />

U<br />

Uganda 1.4<br />

Ultrasonic velocity test 1.11<br />

Unconfined compressive strength 1.5 1.10<br />

Uranium tailings 3.19 11.27<br />

V<br />

Vallentine blockage factor 4.45 4.46<br />

Valves<br />

ball 2.46 2.53<br />

check 2.46 2.51 2.53<br />

coefficient 2.46 2.50<br />

equivalent length 2.46 2.47<br />

foot valve 2.46<br />

knife gate 2.46 2.52 2.53<br />

lift check valve 2.46<br />

pinch 2.53 2.54 2.55<br />

plug 2.53<br />

sw D422 12.21<br />

D698 12.22<br />

D854 12.21<br />

ing check valve 2.46<br />

tilting check valve 2.46<br />

Vedernikov number 6.55 6.56 6.57<br />

Velocity<br />

critical 1.17 6.23 6.26<br />

deposition in o<strong>pe</strong>n channels 6.28 6.29<br />

sinking (see also terminal<br />

below) 1.17 1.18<br />

terminal 3.2 3.3 3.17 4.28


Index terms Links<br />

transition 5.9<br />

Viscous transition 1.19<br />

Vena contracta 2.50<br />

Vertimill 1.30<br />

Viscometer 3.33 3.38<br />

Viscosity<br />

correction for volumetric 1.23<br />

concentration<br />

dynamic 1.15 1.19 1.21 1.22<br />

eddy 6.9<br />

effect on pumps 8.61 8.64<br />

effective in launders 6.31<br />

kinematic 2.4<br />

Viscous sublayer 2.45<br />

Volcanic ash 6.25<br />

Volute 8.2 8.6 8.30<br />

Von Karman constant 4.30 6.9<br />

Vortex flow 8.7 8.8<br />

W<br />

Wall effect on terminal velocity 3.8 3.10<br />

Water physical pro<strong>pe</strong>rties 2.21<br />

Waterfalls 6.58 6.71<br />

Wear 8.4 10.16<br />

Wear plate, pumps 8.2<br />

Weirs, for plunge pools 6.66 6.71<br />

Wenglu Pi<strong>pe</strong>line 1.34<br />

West Irian 1.34<br />

Wet end pump 8.2 8.53<br />

White cast iron 8.3 8.6 10.4 10.11<br />

Wilson and Judge correlation 4.11<br />

Wilson, Snyder pumps 9.7 11.12 11.20<br />

Wilson- Thomas method for<br />

non-Newtonian flows 5.22 5.25 5.30 5.32<br />

Wirth pumps 9.2 9.13<br />

X<br />

Xanthates 7.39<br />

Z<br />

Zandi and Govatos’s equation 4.16 4.21 4.22<br />

Zinc<br />

concentrate 3.19<br />

ore (see also lead, zinc) 7.23


CONTENTS<br />

Preface xvii<br />

PART ONE HYDRAULICS OF SLURRY FLOWS<br />

1 General Concepts of Slurry Flows 1.3<br />

1-0 Introduction 1.4<br />

1-1 Pro<strong>pe</strong>rties of Soils for Slurry Mixtures 1.5<br />

1-1-1 Classifications of Soils for Slurry Mixtures 1.5<br />

1-1-2 Testing of Soils 1.8<br />

1-1-3 Textures of Soils 1.13<br />

1-1-4 Plasticity of Soils 1.13<br />

1-2 Slurry Flows 1.15<br />

1-2-1 Homogeneous Flows 1.16<br />

1-2-2 Heterogeneous Flows 1.16<br />

1-2-3 Intermediate Flow Regimes 1.16<br />

1-2-4 Flows of Emulsions 1.16<br />

1-2-5 Flows of Emulsions - Slurry Mixtures 1.17<br />

1-3 Sinking Velocity of Particles, and Critical Velocity of Flow 1.17<br />

1-3-1 Sinking or Terminal Velocity of Particles 1.17<br />

1-3-2 Critical Velocity of Flows 1.17<br />

1-4 Density of a Slurry Mixture 1.19<br />

1-5 Dynamic Viscosity of a Newtonian Slurry Mixture<br />

1-5-1 Absolute (or Dynamic) Viscosity of Mixtures with Volume<br />

1.21<br />

Concentration Smaller Than 1%<br />

1-5-2 Absolute (or Dynamic) Viscosity of Mixtures with Solids<br />

1.21<br />

with Volume Concentration Smaller Than 20%<br />

1-5-3 Absolute (or Dynamic) Viscosity of Mixtures with High<br />

1.21<br />

Volume Concentration of Solids 1.22<br />

1-6 S<strong>pe</strong>cific Heat 1.22<br />

1-7 Thermal Conductivity and Heat Transfer 1.22<br />

1-8 Slurry Circuits in Extractive Metallurgy 1.24<br />

1-8-1 Crushing 1.24<br />

1-8-2 Milling and Primary Grinding 1.25<br />

1-8-3 Classification 1.26<br />

1-8-4 Concentration and Separation Circuits 1.26<br />

1-8-5 Piping the Concentrate 1.30<br />

1-8-6 Disposal of the Tailings 1.30<br />

1-9 Closed and O<strong>pe</strong>n Channel Flows, Pi<strong>pe</strong>lines Versus Launders 1.31<br />

1-10 Historical Development of Slurry Pi<strong>pe</strong>lines 1.32<br />

vii


viii CONTENTS<br />

1-11 Sedimentation of Dams—A role for the Slurry Engineer 1.33<br />

1-12 Conclusion 1.37<br />

1-13 Nomenclature 1.37<br />

1-14 References 1.38<br />

2 Fundamentals of Water Flows in Pi<strong>pe</strong>s 2.1<br />

2-0 Introduction 2.1<br />

2-1 Shear Stress of Liquid Flows 2.1<br />

2-2 Reynolds Number and Flow Regimes 2.3<br />

2-3 Friction Factors 2.4<br />

2-3-1 Laminar Friction Factors 2.6<br />

2-3-2 Transition Flow Friction Factor 2.8<br />

2-3-3 Friction Factor in Turbulent Flow 2.9<br />

2-3-4 Hazen–Williams Formula 2.18<br />

2.4 The Hydraulic Friction Gradient of Water in Rubber-Lined<br />

Steel Pi<strong>pe</strong>s 2.19<br />

2-5 Dynamics of the Boundary Layer 2.33<br />

2-5-1 Entrance Length 2.33<br />

2-5-2 Friction Velocity 2.35<br />

2-6 Pressure Losses Due to Conduits and Fittings 2.44<br />

2-7 Orifice Plates, Nozzles and Valves Head Losses 2.49<br />

2-8 Pressure Losses Through Fittings at Low Reynolds Number 2.54<br />

2-9 The Bernoulli Equation 2.58<br />

2-10 Energy and Hydraulic Grade Lines with Friction 2.58<br />

2-11 Fundamental Heat Transfer in Pi<strong>pe</strong>s 2.58<br />

2-11-1 Conduction 2.60<br />

2-11-2 Thermal Resistance 2.60<br />

2-11-3 The R Value 2.60<br />

2-11-4 The S<strong>pe</strong>cific Heat or Heat Capacity Cp 2.61<br />

2-11-5 Characteristic Length 2.61<br />

2-11-6 Thermal Diffusivity 2.61<br />

2-11-7 Heat Transfer 2.61<br />

2-12 Conclusion 2.62<br />

2-13 Nomenclature 2.62<br />

2-14 References 2.64<br />

3 Mechanics of Sus<strong>pe</strong>nsion of Solids in Liquids 3.1<br />

3-0 Introduction 3.1<br />

3-1 Drag Coefficient and Terminal Velocity of Sus<strong>pe</strong>nded Spheres<br />

in a Fluid 3.1<br />

3-1-1 The Airplane Analogy 3.1<br />

3-1-2 Buoyancy of Floating Objects 3.3<br />

3-1-3 Terminal Velocity of Spherical Particles<br />

3-1-3-1 Terminal Velocity of a Sphere Falling in a<br />

3.3<br />

Vertical Tube 3.3<br />

3-1-3-2 Very Fine Spheres 3.5<br />

3-1-3-3 Intermediate Spheres 3.6<br />

3-1-3-4 Large spheres 3.7<br />

3-1-4 Effects of Cylindrical Walls on Terminal Velocity 3.8


CONTENTS<br />

3-1-5 Effects of the Volumetric Concentration on the<br />

Terminal Velocity 3.10<br />

3-2 Generalized Drag Coefficient—The Concept of Sha<strong>pe</strong> Factor 3.12<br />

3-3 Non-Newtonian Slurries 3.17<br />

3-4 Time-Inde<strong>pe</strong>ndent Non-Newtonian Mixtures 3.18<br />

3-4-1 Bingham Plastics 3.18<br />

3-4-2 Pseudoplastic Slurries 3.25<br />

3-4-2-1 Homogeneous Pseudoplastics 3.25<br />

3-4-2-2 Pseudohomogeneous Pseudoplastics 3.27<br />

3-4-3 Dilatant Slurries 3.28<br />

3-4-4 Yield Pseudoplastic Slurries 3.28<br />

3-5 Time-De<strong>pe</strong>ndent Non-Newtonian Mixtures 3.30<br />

3-5-1 Thixotropic Mixtures 3.30<br />

3-6 Drag Coefficient of Solids Sus<strong>pe</strong>nded in Non-Newtonian Flows 3.32<br />

3-7 Measurement of Rheology 3.32<br />

3-7-1 The Capillary-Tube Viscometer 3.33<br />

3-7-2 The Coaxial Cylinder Rotary Viscometer 3.36<br />

3-8 Conclusion 3.38<br />

3-9 Nomenclature 3.38<br />

3-10 References 3.41<br />

4 Heterogeneous Flows of Settling Slurries 4.1<br />

4-0 Introduction 4.1<br />

4-1 Regimes of Flow of a Heterogeneous Mixture in Horizontal Pi<strong>pe</strong> 4.2<br />

4-1-1 Flow with a Stationary Bed 4.3<br />

4-1-2 Flow with a Moving Bed 4.3<br />

4-1-3 Sus<strong>pe</strong>nsion Maintained by Turbulence 4.4<br />

4-1-4 Symmetric Flow at High S<strong>pe</strong>ed 4.4<br />

4-2 Hold Up 4.5<br />

4-3 Transitional Velocities 4.5<br />

4-3-1 Transitional Velocities V1 and V2 4.7<br />

4-3-2 The Transitional Velocity V3 or S<strong>pe</strong>ed for Minimum<br />

Pressure Gradient 4.8<br />

4-3-3 V4: Transition S<strong>pe</strong>ed between Heterogeneous and<br />

Pseudohomogeneous Flow 4.18<br />

4-4 Hydraulic Friction Gradient of Horizontal Heterogeneous Flows 4.19<br />

4-4-1 Methods Based on the Drag Coefficient of Particles 4.21<br />

4-4-2 Effect of Lift Forces 4.25<br />

4-4-3 Russian Work on Coarse Coal 4.26<br />

4-4-4 Equations for Nickel–Water Sus<strong>pe</strong>nsions 4.28<br />

4-4-5 Models Based on Terminal Velocity 4.28<br />

4-5 Distribution of Particle Concentration in Compound Systems 4.30<br />

4-6 Friction Losses for Compound Mixtures in Horizontal<br />

Heterogeneous Flows 4.33<br />

4-7 Saltation and Blockage 4.43<br />

4-7-1 Pressure Drop Due to Saltation Flows 4.43<br />

4-7-2 Restarting Pi<strong>pe</strong>lines after Shut-Down or Blockage 4.45<br />

4-8 Pseudohomogeneous or Symmetric Flows 4.47<br />

4-9 Stratified Flows 4.48<br />

4-10 Two-Layer Models 4.50<br />

ix


x CONTENTS<br />

4-11 Vertical Flow of Coarse Particles 4.57<br />

4-12 Inclined Heterogeneous Flows 4.58<br />

4-12-1 Critical Slo<strong>pe</strong> of Inclined Pi<strong>pe</strong>s 4.59<br />

4-12-2 Two-Layer Model for Inclined Flows 4.61<br />

4-13 Conclusion 4.62<br />

4-14 Nomenclature 4.63<br />

4-15 References 4.66<br />

5 Homogeneous Flows of Nonsettling Slurries 5.1<br />

5-0 Introduction 5.1<br />

5-1 Friction Losses for Bingham Plastics 5.2<br />

5-1-1 Start-up Pressure 5.2<br />

5-1-2 Friction Factor in Laminar Regime 5.5<br />

5-1-3 Transition to Turbulent Flow Regime 5.8<br />

5-1-4 Friction Factor in the Turbulent Flow Regime 5.9<br />

5-2 Friction Losses for Pseudoplastics 5.11<br />

5-2-1 Laminar Flow 5.11<br />

5-2-1-1 The Rabinowitsch–Mooney Relations 5.11<br />

5-2-1-2 The Metzner and Reed Approach 5.11<br />

5-2-1-3 The Tomita Method 5.13<br />

5-2-1-3 Heywood Method 5.14<br />

5-2-2 Transition Flow Regime 5.14<br />

5-2-3 Turbulent Flow 5.14<br />

5.3 Friction Losses for Yield Pseudoplastics 5.17<br />

5-3-1 The Hanks and Ricks Method 5.17<br />

5-3-2 The Heywood Method 5.18<br />

5-3-3 The Torrance Method 5.18<br />

5-4 Generalized Methods 5.19<br />

5-4-1 The Hershel–Bulkley Model 5.19<br />

5-4-2 The Chilton and Stainsby Method 5.19<br />

5-4-3 The Wilson–Thomas Method 5.22<br />

5-4-4 The Darby Method: Taking into Account Particle Distribution 5.24<br />

5-5 Time-De<strong>pe</strong>ndent Non-Newtonian Slurries 5.28<br />

5-6 Emulsions 5.29<br />

5-7 Roughness Effects on Friction Coefficients 5.29<br />

5-8 Wall Slippage 5.33<br />

5-9 Pressure Loss through Pi<strong>pe</strong> Fittings 5.34<br />

5-10 Scaling up From Small to Large Pi<strong>pe</strong>s 5.35<br />

5-11 Practical Cases of Non-Newtonian Slurries 5.35<br />

5-11-1 Bauxite Residue 5.35<br />

5-11-2 Kaolin Slurries 5.38<br />

5-12 Drag Reduction 5.39<br />

5-13 Pulp and Pa<strong>pe</strong>r 5.40<br />

5-14 Conclusion 5.41<br />

5-15 Nomenclature 5.42<br />

5-16 References 5.44<br />

6 Slurry Flow In O<strong>pe</strong>n Channels and Drop Boxes 6.1<br />

6-0 Introduction 6.1<br />

6-1 Friction for Single-Phase Flows in O<strong>pe</strong>n Channels 6.2


CONTENTS<br />

6-2 Transportation of Sediments in an O<strong>pe</strong>n Channel 6.9<br />

6-2-1 Measurements of the Concentration of Sediments 6.12<br />

6-2-2 Mean Concentrations for Dilute Mixtures (C v < 0.1) 6.18<br />

6-2-3 Magnitude of � 6.22<br />

6-3 Critical Velocity and Critical Shear Stress 6.23<br />

6-4 Deposition Velocity 6.27<br />

6-5 Flow Resistance and Friction Factor for Heterogeneous Slurry Flows 6.29<br />

6-5-1 Flow Resistances in Terms of Friction Velocity 6.30<br />

6-5-2 Friction Factors 6.31<br />

6-5-2-1 Effect of Roughness 6.31<br />

6-5-2-2 Effect of Particle Concentration on Slurry Viscosity 6.31<br />

6-5-2-3 Effects of Particle Sizes on the Chezy Coefficient 6.32<br />

6-5-2-4 Effect of Bed Form on the Friction 6.33<br />

6-5-3 The Graf–Acaroglu Relation 6.33<br />

6-5-4 Slip of Coarse Materials 6.35<br />

6-5-5 Comparison between Different Models 6.36<br />

6-6 Friction Losses and Slo<strong>pe</strong> for Homogeneous Slurry Flows 6.39<br />

6-6-1 Bingham Plastics 6.40<br />

6-7 Flocculation Launders 6.44<br />

6-8 Froude Number and Stability of Slurry Flows 6.45<br />

6-9 Methodology of Design 6.45<br />

6-10 Slurry Flow in Cascades 6.54<br />

6-11 Hydraulics of the Drop Box and the Plunge Pool 6.56<br />

6-12 Plunge Pools and Drops Followed by Weirs 6.67<br />

6-13 Conclusion 6.71<br />

6-14 Nomenclature 6.71<br />

6-15 References 6.74<br />

PART TWO EQUIPMENT AND PIPELINES<br />

7 Components of Slurry Plants 7.3<br />

7-0 Introduction 7.3<br />

7-1 Rock Crushing 7.3<br />

7-1-1 Primary Crushers 7.4<br />

7-1-1-1 Jaw Crushers 7.5<br />

7-1-1-2 Gyratory Crushers 7.7<br />

7-1-1-3 Impact Crushers 7.8<br />

7-2 Secondary and Tertiary Crushers 7.9<br />

7-2-1 Cone Crushers 7.9<br />

7-2-2 Roll Crushers 7.11<br />

7-3 Grinding Circuits 7.11<br />

7-3-1 Single-Stage Circuits 7.21<br />

7-3-2 Double-Stage Circuits 7.23<br />

7-4 Horizontal Tumbling Mills 7.23<br />

7-4-1 Rod Mills 7.26<br />

7-4-2 Ball Mills 7.26<br />

7-4-3 Autogeneous and Semiautogeneous Mills 7.26<br />

7-5 Agitated Grinding 7.27<br />

7-5-1 Vertical Tower Mills 7.28<br />

xi


xii CONTENTS<br />

7-5-2 Vertical Spindle Mills 7.28<br />

7-5-3 Roller Mills 7.28<br />

7.5.4 Vibrating Ball Mills 7.28<br />

7.5.5 Hammer Mills 7.31<br />

7-6 Screening Devices 7.31<br />

7-6-1 Trommel Screens 7.32<br />

7-6-2 Shaking Screens 7.32<br />

7-6-3 Vibrating Screens 7.32<br />

7-6-4 Banana Screens 7.32<br />

7-7 Slurry Classifiers 7.32<br />

7-7-1 Hydraulic Classifiers 7.32<br />

7-7-2 Mechanical Classifiers 7.33<br />

7-7-3 Hydrocyclones 7.33<br />

7-7-4 Magnetic Separators 7.38<br />

7-8 Flotation Circuits 7.38<br />

7-9 Mixers and Agitators 7.40<br />

7-10 Sedimentation 7.59<br />

7-10-1 Gravity Sedimentation 7.60<br />

7-10-2 Centrifuges 7.62<br />

7-11 Conclusion 7.64<br />

7-12 Nomenclature 7.64<br />

7-13 References 7.66<br />

8 The Design of Centrifugal Slurry Pumps 8.1<br />

8.0 Introduction 8.1<br />

8.1 The Centrifugal Slurry Pump 8.2<br />

8.2 Elementary Hydraulics of the Slurry Pump 8.6<br />

8.2.1 Vortex Flow 8.7<br />

8-2-2 The Ideal Euler Head 8.8<br />

8-2-3 Slip of Flow Through Im<strong>pe</strong>ller Channels 8.11<br />

8-2-4 The S<strong>pe</strong>cific S<strong>pe</strong>ed 8.14<br />

8-2-5 Net Positive Suction Head and Cavitation 8.18<br />

8-3 The Pump Casing 8.25<br />

8-4 The Im<strong>pe</strong>ller, the Ex<strong>pe</strong>ller and the Dynamic Seal 8.34<br />

8-5 Design of the Drive End 8.42<br />

8-5-1 The Radial Thrust Due To Total Dynamic Head 8.43<br />

8-5-2 The Axial Thrust Due to Pressure 8.43<br />

8-5-3 Thread Pull Force 8.48<br />

8-5-4 Radial Force on the Drive End 8.51<br />

8-5-5 Total Forces from the Wet End 8.51<br />

8-5-6 Flange Loads 8.52<br />

8-6 Adjustment of the Wet End 8.53<br />

8-7 Vertical Slurry Pumps 8.53<br />

8-8 Gravel and Dredge Pumps 8.59<br />

8-9 Affinity Laws 8.60<br />

8-10 Performance Corrections for Slurry Pumps 8.61<br />

8-10-1 Corrections for Viscosity and Slip<br />

8-10-2 Concepts of Head Ratio and Efficiency Ratio Due to<br />

8.61<br />

Pumping Solids 8.64


CONTENTS<br />

8-10-3 Concepts of Head Ratio and Efficiency Ratio Due to<br />

Pumping Froth 8.68<br />

8-11 Conclusion 8.72<br />

8-12 Nomenclature 8.72<br />

8-13 References 8.75<br />

9 Positive Displacement Pumps 9.1<br />

9-0 Introduction 9.1<br />

9-1 Solid Piston Pumps 9.1<br />

9-2 Plunger Pumps 9.6<br />

9-3 Diaphragm Piston Pumps 9.8<br />

9-4 Accessories for Piston and Plunger Pumps 9.13<br />

9-5 Peristaltic Pumps 9.13<br />

9-6 Rotary Lobe Slurry Pumps 9.14<br />

9-7 The Lockhop<strong>pe</strong>r Pump 9.15<br />

9-8 Conclusion 9.16<br />

9-9 References 9.17<br />

10 Materials Science for Slurry Systems 10.1<br />

10.0 Introduction 10.1<br />

10-1 The Stress- Strain Relationship of Metals 10.1<br />

10-2 Iron and Its Alloys for the Slurry Industry 10.3<br />

10-2-1 Grey Iron 10.3<br />

10-2-2 Ductile Iron 10.4<br />

10.3 White Iron 10.4<br />

10-3-1 Malleable Iron 10.4<br />

10-3-2 Low-Alloy White Irons 10.5<br />

10-3-3 Ni-Hard 10.5<br />

10-3-4 High-Chrome–Molybdenum Alloys 10.6<br />

10.4 Natural Rubbers 10.11<br />

10-4-1 Natural Aashto 10.12<br />

10-4-2 Pure Tan Gum 10.12<br />

10-4-3 White Food-Grade Natural Rubber 10.12<br />

10-4-4 Carbon-Black-Filled Natural Rubber 10.13<br />

10-4-5 Carbon-Black- and Silicon-Filled Natural Rubber<br />

10-4-6 Hard Natural Rubber/ Butadiene Styrene Compound<br />

10.13<br />

Filled with Graphite 10.13<br />

10-5 Synthetic Rubbers 10.13<br />

10-5-1 Polychlorene (Neoprene) 10.14<br />

10-5-2 Ethylene Propylene Terpolymer (EPDM) 10.15<br />

10-5-3 Jade Green Armabond 10.15<br />

10-5-4 Armadillo 10.15<br />

10-5-5 Nitrile 10.15<br />

10-5-6 Carboxylic Nitrile 10.17<br />

10-5-7 Hypalon 10.17<br />

10-5-8 Fluoro-elastomer (Viton) 10.18<br />

10-5-9 Polyurethane 10.18<br />

10-6 Wear Due to Slurries 10.18<br />

10-7 Conclusion 10.21<br />

10-8 References 10.22<br />

xiii


xiv CONTENTS<br />

11 Slurry Pi<strong>pe</strong>lines 11.1<br />

11.0 Introduction 11.1<br />

11-1 Bauxite Pumping 11.1<br />

11-2 Gold Tailings 11.2<br />

11-3 Coal Slurries 11.2<br />

11-3-1 Size of Coal Particles 11.2<br />

11-3-2 Degradation of Coal During Hydraulic Transport 11.3<br />

11-3-3 Coal–Magnetite Mixtures 11.4<br />

11-3-4 Chemical Additions to Coal–Water Mixtures. 11.5<br />

11-3-5 Coal–Oil Mixtures 11.5<br />

11-3-6 Dewatering Coal Slurry 11.6<br />

11-3-7 Ship Loading Coarse Coal 11.8<br />

11-3-8 Combustion of Coal–Water Mixtures (CWM) 11.8<br />

11-3-9 Pumping Coal Slurry Mixtures 11.10<br />

11-4 Limestone Pi<strong>pe</strong>lines 11.10<br />

11-5 Iron Ore Slurry Pi<strong>pe</strong>lines 11.12<br />

11-6 Phosphate and Phosphoric Acid Slurries 11.16<br />

11-6-1 Rheology 11.17<br />

11-6-2 Materials Selection for Phosphate 11.18<br />

11-6-3 The Chevron Pi<strong>pe</strong>line 11.19<br />

11-6-4 The Goiasfertil Phosphate Pi<strong>pe</strong>line 11.20<br />

11-6-5 The Hindustan Zinc Phosphate Pi<strong>pe</strong>line 11.21<br />

11-7 Cop<strong>pe</strong>r Slurry and Concentrate Pi<strong>pe</strong>lines 11.21<br />

11-8 Clay and Drilling Muds 11.22<br />

11-9 Oil Sands 11.23<br />

11-10 Backfill Pi<strong>pe</strong>lines 11.24<br />

11-11 Uranium Tailings 11.27<br />

11-12 Codes and Standards for Slurry Pi<strong>pe</strong>lines 11.27<br />

11-13 Conclusion 11.30<br />

11-14 References 11.31<br />

12 Feasibility Study for A Slurry Pi<strong>pe</strong>line<br />

and Tailings Disposal System 12.1<br />

12-0 Introduction 12.1<br />

12-1 Project Definition 12.2<br />

12-2 Rheology, Thickeners Performance, Pi<strong>pe</strong>line Sizing 12.5<br />

12-3 Reclaim Water Pi<strong>pe</strong>line 12.8<br />

12-4 Emergency Pond 12.9<br />

12-5 Tailings Dams 12.11<br />

12-5-1 Wall Building by Spigotting 12.11<br />

12-5-2 Deposition by Cycloning 12.12<br />

12-5-2-1 Mobile Cycloning by the Upstream Method 12.14<br />

12-5-2-2 Mobile Cycloning by the Downstream Method 12.14<br />

12-5-2-3 Deposition by Centerline 12.15<br />

12-5-2-4 Multicellular Construction 12.15<br />

12-6 Submerged Disposal 12.15<br />

12-6-1 Subsea Deposition Techniques 12.17<br />

12-7 Tailings Dam Design 12.17<br />

12-8 Seepage Analysis of Tailings Dams 12.18


CONTENTS<br />

12-9 Stability Analysis for Tailings Dams 12.18<br />

12-10 Erosion and Corrosion 12.19<br />

12-11 Hydraulics 12.19<br />

12-12 Pump Station Design 12.19<br />

12-13 Electric Power System 12.20<br />

12-14 Telecommunications 12.21<br />

12-15 Tailings Dam Monitoring 12.21<br />

12-16 Choke Stations and Impactors 12.22<br />

12-17 Establishing an Approach for Start-up and Shutdown 12.22<br />

12-18 Closure and Reclamation Plan 12.23<br />

12-19 Access and Service Roads 12.24<br />

12-20 Cost Estimates 12.24<br />

12-20-1 Capital Costs 12.24<br />

12-20-2 O<strong>pe</strong>ration Cost Estimates 12.25<br />

12-21 Project Implementation Plan 12.27<br />

12-22 Conclusion 12.27<br />

12-23 References 12.28<br />

Ap<strong>pe</strong>ndix A S<strong>pe</strong>cific Gravity and Hardness of Minerals A.1<br />

Ap<strong>pe</strong>ndix B Units of Measurement B.1<br />

Index I.1<br />

xv


PART ONE<br />

HYDRAULICS OF<br />

SLURRY FLOWS


CHAPTER 1<br />

GENERAL CONCEPTS OF<br />

SLURRY FLOWS<br />

1-0 INTRODUCTION<br />

Slurry is essentially a mixture of solids and liquids. Its physical characteristics are de<strong>pe</strong>ndent<br />

on many factors such as size and distribution of particles, concentration of<br />

solids in the liquid phase, size of the conduit, level of turbulence, tem<strong>pe</strong>rature, and<br />

absolute (or dynamic) viscosity of the carrier. Nature offers examples of <strong>slurry</strong> flows<br />

such as seasonal floods that carry silt and gravel. Every year during the flood season,<br />

the Nile transports massive amounts of silt over thousands of miles to the Saharan<br />

desert. To rephrase Herodotus, who once said “Egypt is the gift of the Nile,” one may<br />

consider that one of the most ancient civilizations was de<strong>pe</strong>ndent on natural <strong>slurry</strong> flows<br />

for its survival.<br />

Dredging is one of the most common and ancient processes involving <strong>slurry</strong> flows; the<br />

dredged materials contain a wide range of particles, tree debris, rocks, etc. Mining has<br />

employed the concept of <strong>slurry</strong> flows in pi<strong>pe</strong>lines since the mid-nineteenth century, when<br />

the technique was used to reclaim gold from placers in California. Long-distance <strong>slurry</strong><br />

pi<strong>pe</strong>lines have evolved in all continents since the mid 1950s. Some <strong>slurry</strong> mixtures consist<br />

of very fine solids at high concentration, such as those in the cop<strong>pe</strong>r concentrate<br />

pi<strong>pe</strong>lines of Escondida, Chile, and Bajo Alumbrera, Argentina. Other mixtures are based<br />

on coarse particles up to a size of 150 mm (6�), such as those pum<strong>pe</strong>d from fields of phosphate<br />

matrix.<br />

This chapter introduces some of the basic principles of <strong>slurry</strong> mixtures and flows. The<br />

<strong>slurry</strong> engineer has to appreciate the pro<strong>pe</strong>rties of the soil to be mined, dredged, or mixed<br />

with water. Original rock sizes, hardness, and plasticity play a major role in the selection<br />

of the equipment for crushing, milling, flotation, tailings disposal, or soil reclamation.<br />

Understanding sinking and critical s<strong>pe</strong>eds are essential when sizing the pi<strong>pe</strong>line. A brief<br />

introduction to <strong>slurry</strong> flows in extractive metallurgy serves the purpose of focusing on the<br />

essentials of the application of <strong>slurry</strong> flows to engineering.<br />

Natural <strong>slurry</strong> flows, even in very dilute forms, can have negative effects on the environment<br />

if not pro<strong>pe</strong>rly managed. Some of the great dams of the world built in the<br />

twentieth century are starting to suffer from siltation. Behind such dams, large lakes are<br />

often man-made. The river flow is brought to a sufficiently slow s<strong>pe</strong>ed for the silt to deposit<br />

at the bottom. Engineers in the twenty-first century will have to learn to manage<br />

the siltation of large man-made lakes using the science of dredging and piping <strong>slurry</strong><br />

flows.<br />

1.3


1.4 CHAPTER ONE<br />

1-1 PROPERTIES OF SOILS FOR<br />

SLURRY MIXTURES<br />

Slurry flows occur in nature in different ways. They are often associated with the transportation<br />

of silt from one region to another. Strong rains lead to soil erosion, mud slides,<br />

and the eventual drainage of slurries toward rivers. These are dilute slurries, in the sense<br />

that the soils mix naturally at a weight ratio of solids to liquids smaller than 15%.<br />

One very interesting river is the Nile. It may be said that during two months of the<br />

year it becomes a massive <strong>slurry</strong> flow. Torrential tropical rains over Lake Victoria in<br />

Uganda and Kenya are the source of the White Nile. Torrential tropical rains over the<br />

Ethiopian plateau are the source of the Blue Nile. On their way to the Sudan, both<br />

branches of this longest river in the world transport silt and soils. The White Nile seems<br />

to lose a lot of its water as it enters the swamps of the Bahr El Ghazal in Sudan. What<br />

is left of the White Nile joins the Blue Nile near Khartoum in Sudan. The Nile pursues<br />

its trip to the north and gradually enters the Saharan desert through Nubia and Egypt. As<br />

the flood season terminates, the silt transported by the Nile sediments by gravity. The<br />

silt has deposited for thousands of years, creating a narrow strip of rich farmland. Out<br />

of this silt grew the towns and states in Nubia and Egypt. The Pharaohs built an advanced<br />

civilization on the silt brought to them by the Nile’s natural <strong>slurry</strong> flows. The<br />

“gift of the Nile” was silt that would not have been deposited without a form of natural<br />

<strong>slurry</strong> flow.<br />

A simplified flow sheet (Figure 1-1) of the Nile illustrates this natural <strong>slurry</strong> flow. The<br />

steps in the process are:<br />

� Water from the rains is the carrier liquid.<br />

� The flow of water from the mountains of Uganda and Kenya moves fast enough during<br />

the flood season to scour the ground of silt and transport it in the form of a dilute <strong>slurry</strong>.<br />

(This is a step of <strong>slurry</strong> formation.)<br />

Uganda/Kenya<br />

Ethiopia<br />

floods<br />

floods<br />

torential<br />

rains<br />

rains<br />

silt transported<br />

by the White Nile<br />

silt transported<br />

by the Blue Nile<br />

Sedimentation<br />

at Bahr El Ghazal<br />

Sudan<br />

Nubia<br />

The Saharan Desert<br />

Egypt<br />

sedimentation by gravity<br />

of the silt after the flood<br />

(Egypt is the Gift of the Nile)<br />

FIGURE 1-1 There is no better example of the importance of <strong>slurry</strong> to civilization than the<br />

land of Egypt. For thousands of years, the Nile has transported massive quantities of silt over<br />

thousands of kilometers to cover by its floods a narrow stretch of land. From these silt layers, a<br />

civilization grew.


GENERAL CONCEPTS OF SLURRY FLOWS<br />

� As the waters from the rains over the mountains of Uganda and Kenya join, they form<br />

the White Nile. (This step is natural hydrotransport.)<br />

� As the White Nile enters the Bahr El Ghazal in Sudan, it spreads and stagnates, forming<br />

swamps. A nomadic life has long flourished around these swamps. (This step involves<br />

partial sedimentation by stagnation in the swamps.)<br />

� In another region (in Ethiopia), rains form the Blue Nile. The flow of water from the<br />

mountains of Ethiopia move fast enough during the flood season to scour the ground of<br />

silt and transport it in the form of a dilute <strong>slurry</strong>. (This is another step of <strong>slurry</strong> formation.)<br />

� The Blue and White Nile merge near Khartoum, Sudan, and continue their flow to the<br />

north.<br />

� As the floods enter Nubia and Egypt, they overflow the banks of the Nile and transport<br />

s<strong>pe</strong>ed of the <strong>slurry</strong> mixture drops.<br />

� Sedimentation of silt occurs, with Egypt acting as a massive clarifier for the waters of<br />

the Nile, particularly at its delta with the Mediterranean Sea. (This step is natural gravity<br />

sedimentation.)<br />

For thousands of years the Pyramids and the Sphinx have stared at this immense natural<br />

<strong>slurry</strong> clarifier that is the Valley of the Nile in the middle of the Saharan Desert (Figure 1-<br />

2).<br />

Dredging is an important engineering activity in which gravel is moved in the form of<br />

<strong>slurry</strong> into a hop<strong>pe</strong>r on a s<strong>pe</strong>cially constructed boat (Figure 1-4). A s<strong>pe</strong>cial pump is often<br />

used in a drag arm (Figure 1-3), and a s<strong>pe</strong>cial suction mouthpiece (Figure 1-5) is used at<br />

the tip of the drag arm.<br />

To complete dredging and form the <strong>slurry</strong>, it is essential to cut through the sand layers,<br />

rocks, and debris, using s<strong>pe</strong>cial cutters for sand (Figure 1-6a) and for rocks (Figure 1-6b)<br />

with very hard, replaceable blades.<br />

The composition of a <strong>slurry</strong> mixture de<strong>pe</strong>nds on many factors such as particle size and<br />

distribution. Particles may be found in nature as soils or may be created by the processes<br />

of crushing, milling, and grinding. For applications such as dredging, natural soils are<br />

pum<strong>pe</strong>d without any crushing or grinding. For mining processes, an understanding of the<br />

physical pro<strong>pe</strong>rties of soils is essential for sizing equipment, crushing and milling, <strong>slurry</strong><br />

preparation, mixing, and pumping (see Figure 1-7).<br />

1-1-1 Classifications of Soils for Slurry Mixtures<br />

There are a variety of methods used to classify soils. Two main classes are:<br />

1. Cohesive soils such as certain silts and clays with a median particle diameter smaller<br />

than 0.0625 mm (less than 0.0025 in, or mesh 250)<br />

2. Noncohesive soils such as certain silts and clays with a median particle diameter larger<br />

than 0.0625 mm (larger than 0.0025 in, or mesh 250)<br />

For underwater dredging, the rock’s strength is determined by its core, and this pro<strong>pe</strong>rty<br />

has a very important effect on the efficiency of dredging. Herbrich (1991) proposed a<br />

classification of soils in terms of unconfined compressive strength (see Table 1-1).<br />

The Permanent International Association of Navigation Congresses (1972) adopted a<br />

system of classification of soils, reviewed by Sargent (1984) and summarized in Tables<br />

1.5


1.6 CHAPTER ONE<br />

FIGURE 1-2 For five thousand years, the Sphinx and the Pyramids have stared from the<br />

Gizeh plateau in the desert at history and at the Nile, which transforms itself every summer into<br />

a natural <strong>slurry</strong> transporter, bringing silt and life to the desert.<br />

1-2, 1-3, and 1-4, that is recommended for use in dredging. In these tables, visual ins<strong>pe</strong>ction<br />

is mentioned as a quick way to determine the nature of soils. This method does not<br />

relieve the engineer from the responsibility of conducting a pro<strong>pe</strong>r size distribution test<br />

and rheology test before any design.<br />

The Standard D2488 of the American Society for Testing of Materials (ASTM)<br />

(1993) also offers a classification of soils, with a range of particle sizes as presented in<br />

Table 1-5. This standard is widely used in North America.


GENERAL CONCEPTS OF SLURRY FLOWS<br />

discharge pi<strong>pe</strong><br />

electric cable<br />

pump<br />

bottom of lake<br />

drag arm column<br />

FIGURE 1-3 Dredging boat and dredge arm.<br />

FIGURE 1-4 S<strong>pe</strong>cial dredger boat.<br />

hop<strong>pe</strong>r for solids<br />

1.7


1.8 CHAPTER ONE<br />

FIGURE 1-5 Suction mouthpiece for boat dredger. Courtesy of Mobile Pulley and Machine<br />

Works.<br />

1-1-2 Testing of Soils<br />

Various soil tests are recommended before mixing the soil with water in the early stages<br />

of designing a dredging or <strong>slurry</strong> transportation system. Particle size distribution should<br />

be established. Table 1-6 presents conversion factors between the three most common<br />

scales for measuring particle size.<br />

A number of tests are recommended to determine the dredgeability of soils and their<br />

behavior in placer mining or <strong>slurry</strong> mixing (Table 1-7). In nature, silts may be found in<br />

association with clays; thus, the parameters for both silts and clays should be assessed.<br />

The following testing parameters are accepted by the industry.<br />

Composition Tests<br />

� Visual ins<strong>pe</strong>ction: For the purpose of assessment of the rock mass. Such a test indicates<br />

the in situ state of the rock mass. Tests may be conducted in situ or under lab<br />

conditions in accordance with British Standard Institute Standard BS 5930 (1999).<br />

� Section thickness test: A lab test conducted for the purpose of geotechnical identification<br />

and as a tool to determine mineral composition of the rock mass.<br />

� Bulk density: Wet and dry tests are conducted under laboratory conditions to assess the<br />

weight and volume relationship. (International Journal of Rock Mechanics and Mineral<br />

Sciences, 1979).<br />

� Porosity: This is a calculation of voids as a <strong>pe</strong>rcentage of total volume and is based on<br />

lab tests on bulk density.<br />

� Carbonate content: This lab test should be conducted in accordance with American Society<br />

for Testing Materials (ASTM) Standard D3155 (1983) to measure lime content,<br />

particularly in limestone and chalks.


(a)<br />

(b)<br />

FIGURE 1-6 (a) S<strong>pe</strong>cial dredging sand cutter. The blades are replaceable. Courtesy of Mobile<br />

Pulley and Machine Works. (b) S<strong>pe</strong>cial dredging rock cutter. Courtesy of Mobile Pulley<br />

and Machine Works.<br />

1.9


1.10 CHAPTER ONE<br />

FIGURE 1-7 Mineral process plants can reject fairly coarse material that is left after crushing<br />

and milling mineral rocks. In this case, the coarse material is transported by piping in the<br />

form of a tailings <strong>slurry</strong> and used to build a tailings dam.<br />

Strength, Hardness, and Stratification Tests<br />

� Surface hardness: This lab test should be conducted to determine hardness in terms of<br />

the Mohr’s scale (from 0 for talc to 10 for diamonds). Ap<strong>pe</strong>ndix I presents a tabulation<br />

of density and Mohr hardness of minerals. The hardness of minerals is critical to the<br />

wear life of equipment associated with <strong>slurry</strong> flows.<br />

� Uniaxial compression: This lab test measures ultimate strength under uniaxial stress.<br />

These tests should be done on fully saturated samples. The dimensions of the test sample<br />

and the directions of stratification influence stress direction. Cylinder samples<br />

TABLE 1-1 Classification of Soils in Terms of Unconfined Compressive Strength.<br />

(After Herbrich, 1991)<br />

Unconfined compressive strength<br />

Characteristic MPa 10 3 psi<br />

Very weak < 1.25 < 0.145<br />

Weak 1.25–5.0 0.15–0.73<br />

Moderately weak 5.0–12.5 0.73–1.8<br />

Moderately strong 12.5–50.0 1.8–7.3<br />

Strong 50–100 7.3–14.6<br />

Very strong 100–200 14.6–29.2<br />

Extremely strong > 200 > 29.2


GENERAL CONCEPTS OF SLURRY FLOWS<br />

TABLE 1-2 Classification of Noncohesive Dredged Soils after the Permanent<br />

International Association of Navigation Congresses (1972, 1984)<br />

should have a length-to-diameter ratio of 2:1, as <strong>pe</strong>r The International Society for Rock<br />

Mechanics (1978).<br />

� Brazilian split: This is a lab test to measure strength as derived from uniaxial testing.<br />

This procedure is similar to the uniaxial compression test but with a different lengthto-diameter<br />

ratio. For further details, consult The International Society for Rock Mechanics<br />

(1977).<br />

� Point load test: This is a quick lab test to measure strength. It should be conducted with<br />

the uniaxial compression test as described by Broch and Franklin (1972).<br />

� Seismic velocity test: This field in situ test is conducted to check on the stratigraphy<br />

and fracturing of rock masses. It is useful for extrapolating field and lab measurements<br />

to rock mass behavior.<br />

� Ultrasonic velocity test: This lab test is conducted on cores in the longitudinal direction.<br />

� Static modulus of elasticity: This lab test measures stress/strain rate and gives an indication<br />

of the brittleness of rock.<br />

� Drillability: This in situ test measures <strong>pe</strong>netration rate, torque, feed force, fluid pressure,<br />

depth of layers, etc., and is used to establish the drill techniques and s<strong>pe</strong>cification<br />

for placer mining or dredging.<br />

� Angularity: This lab test is conducted to assess the sha<strong>pe</strong> of particles by visual ins<strong>pe</strong>ction<br />

in accordance with British Standard Institute BS 812 (1999).<br />

The ex<strong>pe</strong>rtise of a geologist is essential for mining or dredging large areas.<br />

1.11<br />

Ty<strong>pe</strong><br />

Identification of particle sizes<br />

Strength and<br />

of soils mm BS sieve units Identification structural pro<strong>pe</strong>rties<br />

Boulders > 200 6 Visual<br />

and 60–200 examination and<br />

cobbles measurement<br />

Gravel Fine 2–6 mm Fine No. 7— 1 Medium 6–20 mm<br />

–<br />

4 in<br />

Medium<br />

Visual May be found loose<br />

1 –<br />

4– 3 Coarse 20–80 mm<br />

–<br />

4 in<br />

Coarse<br />

examination in some fields, or in<br />

3 –<br />

4–3 in cemented beds, or<br />

may ap<strong>pe</strong>ar as weak<br />

conglomerate beds or<br />

hard packed gravel<br />

intermixed with sand<br />

Sands Fine 0.06–0.2 mm Fine mesh 72–200 Visual Strength varies<br />

Medium 0.2–0.6 mm Medium mesh 25–72 examination. No between compacted,<br />

Coarse 0.6–2 mm Coarse mesh 7–25 cohesion when loose and cemented.<br />

dry Homogeneous or<br />

stratified structures.<br />

Intermixture with silt<br />

or clay may produce<br />

hardpacked sands


TABLE 1-3 Classification of Cohesive Natural Soils after the Permanent International<br />

Association of Navigation Congresses (1972, 1984)<br />

Ty<strong>pe</strong><br />

Identification of particle sizes<br />

Strength and<br />

of soils mm BS sieve Identification structural pro<strong>pe</strong>rties<br />

Silts Fine 0.002–0.006 Passing No. Individual particles Coarse and sandy particles are<br />

Medium 0.006–0.02 200 are invisible. Wet nonplastic but similar<br />

Coarse 0.02–0.06 lumps or coarse are characteristics to sands. Fine<br />

visible. Determination silts are plastic and similar to<br />

by testing for clays. They are often found in<br />

dilatancy*. Silt can nature intermixed with sand<br />

be dusted off fingers and clay. They may be homoafter<br />

drying and dry geneous or stratified and their<br />

lumps are powdered consistency may vary from<br />

by finger pressure fluid silt to stiff silt or siltstone<br />

Clays Finer than 0.002 N/A Clays are very<br />

Strength Shear strength<br />

cohesive and are Very soft: may < 20kN/m2 plastic without be squeezed < 2.9 psi<br />

dilatancy. Moist easily between<br />

samples stick to fingers<br />

fingers with smooth,<br />

Soft: easily 20–40 kN/m2 greasy touch. Dry<br />

lumps do not powder.<br />

molded by<br />

fingers<br />

2.9–5.8 psi<br />

Clays shrink and Firm: requires 40–75 kN/m2 crack by drying and strong pressure 5.8–10.9 psi<br />

develop high strength to mold by<br />

fingers<br />

Structure of clays may Stiff: can not 75–150 kN/m2 be fissured, intact, be molded by 10.9–21.8 psi<br />

homogeneous, fingers, dent<br />

stratified, or by thumbnail<br />

weathered.<br />

Hard: tough, Above 150<br />

intended with kN/m2 difficulty by<br />

thumbnail<br />

21.8 psi<br />

*Dilatancy is a pro<strong>pe</strong>rty exhibited by silt when shaken, and is due to high <strong>pe</strong>rmeability of silt. When<br />

a moistened sample is shaken in the o<strong>pe</strong>n hand, water ap<strong>pe</strong>ars on the surface, giving it a glossy ap<strong>pe</strong>arance.<br />

TABLE 1-4 Classification of Organic Soils after the Permanent International<br />

Association of Navigation Congresses (1972, 1984)<br />

Ty<strong>pe</strong><br />

Identification of particle sizes<br />

Strength and<br />

of soils mm BS sieve Identification structural pro<strong>pe</strong>rties<br />

Peat and N/A N/A It is generally identified It may be firm or spongy in<br />

organic as brown or black with a nature and its strength is<br />

soils strong organic smell and different in horizontal and<br />

contains wood and fibers. vertical directions.<br />

1.12


1-1-3 Textures of Soils<br />

Granular soils are found in nature as a mixture of particles of different sizes. Two coefficients<br />

are used to express such texture:<br />

1. The coefficient of curvature, C c (equation 1-1)<br />

2. The coefficient of uniformity, C u (equation 1-2)<br />

C c = (1-1)<br />

Cu = � (1-2)<br />

D10<br />

Where D10, D30, and D60 are defined as the grain size at which 10%, 30%, and 60% of the<br />

soil is finer. According to Herbrich (1991)<br />

If 1 < C c < 3, the grain size distribution will be smooth<br />

If C u > 4 for gravels then there is a wide range of sizes<br />

If C u > 6 for sands then there is a wide range of sizes<br />

Alternatively, the soil is said to contain very little fines and is well graded.<br />

1-1-4 Plasticity of Soils<br />

GENERAL CONCEPTS OF SLURRY FLOWS<br />

D2 30<br />

�<br />

(D60D10) D 60<br />

1.13<br />

For clays and silts, an additional test for the liquid limit (L L) and the plastic limit (P L) are<br />

recommended.<br />

The liquid limit is defined as the moisture content in soil above which it starts to act as<br />

a liquid and below which it acts as a plastic. To conduct a test, a sample of clay is thoroughly<br />

mixed with water in a brass cup. The number of bumps required to close a groove<br />

cut in the pot of clay in the cup is then measured. This test is called the Atterberg test.<br />

The plastic limit is defined as the limit below which the clay will stop behaving as a<br />

plastic and will start to crumble. To measure such a limit, a sample of the soil is formed<br />

into a tubular sha<strong>pe</strong> with a diameter of 3.2 mm (0.125 in) and the water content is measured<br />

when the cylinder ceases to roll and becomes friable.<br />

TABLE 1-5 Range of Particle Sizes of Soils According to ASTM D2488 (1993)<br />

Material Range of sizes in mm Range of sizes in inches<br />

Boulders > 300 > 12<br />

Cobbles 75–300 3–12<br />

Coarse gravel 19–75 0.75–3<br />

Fine gravel 4.75–19 0.019–0.75<br />

Coarse sand 2.00–4.75 0.08–0.0188<br />

Medium sand 0.43–2.00 0.017–0.08<br />

Fine sand 0.08–0.43 0.003–0.017<br />

Silts and clays < 0.075 < 0.003


TABLE 1-6 Conversion between Scales of Particle Size<br />

Sieve o<strong>pe</strong>ning Sieve o<strong>pe</strong>ning<br />

U.S. no. Tyler mesh (micrometers) (inches) Grade of soils<br />

3<br />

2<br />

1.50<br />

Screen shingle gravel<br />

26670 1.050<br />

22430 0.883<br />

18850 0.742<br />

15850 0.624<br />

13330 0.525<br />

11200 0.441<br />

9423 0.371<br />

2.5 2.5 7925 0.321<br />

3 3 6680 0.263<br />

3.5 3.5 5613 0.221<br />

4 4 4699 0.185<br />

5 5 3962 0.156<br />

6 6 3327 0.131<br />

7 7 2794 0.110<br />

8 8 2362 0.093 Very coarse sand<br />

9 9 1981 0.078<br />

10 10 1651 0.065<br />

12 12 1397 0.055<br />

14 14 1168 0.046 Coarse sand<br />

16 16 991 0.039<br />

20 20 833 0.0328<br />

24 24 701 0.0276<br />

28 28 589 0.0232 Medium sand<br />

32 32 495 0.0195<br />

35 35 417 0.0164<br />

42 42 351 0.0138<br />

50 50 297 0.0117<br />

60 60 250 0.01 Fine sand<br />

70 70 210 0.0823<br />

80 80 177 0.07<br />

100 100 149 0.06<br />

120 120 125 0.05<br />

140 140 105 0.041<br />

170 170 88 0.034 Silt<br />

200 200 74 0.029<br />

230 250 63 0.025<br />

270 53 0.02<br />

325 43 0.017 Pulverized silt<br />

400 38 0.015<br />

500 25 0.01<br />

625 20 0.008<br />

1250 10 0.004<br />

2500 5 0.002<br />

12500 1 0.0004<br />


The difference between the liquid and plastic limits is defined as the plasticity index:<br />

PI = LL – PL (1-3)<br />

1-2 SLURRY FLOWS<br />

GENERAL CONCEPTS OF SLURRY FLOWS<br />

TABLE 1-7 Testing Parameters on Soils to Determine<br />

Dredgeability, Suitability for Placer Mining, or Slurry Preparation<br />

Ty<strong>pe</strong> of soil Testing parameters<br />

Sand Density<br />

Water content<br />

S<strong>pe</strong>cific gravity of grains<br />

Grain size<br />

Water <strong>pe</strong>rmeability<br />

Frictional pro<strong>pe</strong>rties<br />

Lime content<br />

Organic content<br />

Silt Density<br />

Water content<br />

Water <strong>pe</strong>rmeability<br />

Shear strength or sliding resistance<br />

Plasticity<br />

Lime content<br />

Organic content<br />

Clay Density<br />

Water content<br />

Sliding resistance<br />

Consistency ranges (plasticity)<br />

Organic content<br />

Peat Same parameters as clay<br />

Gravel Same parameters as sand<br />

1.15<br />

A <strong>slurry</strong> mixture is essentially a mixture of a carrying fluid and solid particles held in sus<strong>pe</strong>nsion.<br />

The most commonly used fluid is water, but over the years, attempts have been<br />

made to use crude oils with milled coal, and even air in pneumatic conveying.<br />

The flow of <strong>slurry</strong> in a pi<strong>pe</strong>line is much different from the flow of a single-phase liquid.<br />

Theoretically, a single-phase liquid of low absolute (or dynamic) viscosity can be allowed<br />

to flow at slow s<strong>pe</strong>eds from a laminar flow to a turbulent flow. However, a twophase<br />

mixture, such as <strong>slurry</strong>, must overcome a deposition critical velocity or a viscous<br />

transition critical velocity. The analogy can be made here in terms of an airplane: if the<br />

s<strong>pe</strong>ed drops excessively, the airplane stalls and stops flying. If the <strong>slurry</strong>’s s<strong>pe</strong>ed of flow<br />

is not sufficiently high, the particles will not be maintained in sus<strong>pe</strong>nsion. On the other<br />

hand, in the case of highly viscous mixtures, if the shear rate in the pi<strong>pe</strong>line is excessively<br />

low, the mixture will be too viscous and will resist flow.<br />

Sections 1-2-1 and 1-2-2 define the two basic <strong>slurry</strong> flows.


1.16 CHAPTER ONE<br />

1-2-1 Homogeneous Flows<br />

Solids are uniformly distributed throughout the liquid carrier. An example of homogeneous<br />

flows is cop<strong>pe</strong>r concentrate <strong>slurry</strong> after undergoing a process of grinding and thickening.<br />

Particles are then very fine and the mixture is at a high concentration (50–60% by<br />

weight). As the concentration of particles is increased (beyond 40% by weight for many<br />

slurries), the mixture becomes more viscous and develops non-Newtonian pro<strong>pe</strong>rties.<br />

Apart from rich concentrate slurries, drilling mud, sewage sludge, and fine limestone (cement<br />

kiln feed <strong>slurry</strong>) behave as homogeneous flows. Typical particle sizes for homogeneous<br />

mixtures are smaller than 40 �m to 70 �m (325–200 mesh), de<strong>pe</strong>nding on the density<br />

of the solids.<br />

The presence of clays in certain circuits must not be ignored. If clay is not separated,<br />

the <strong>slurry</strong> can be quite viscous. Pumps and pi<strong>pe</strong>s must be sized pro<strong>pe</strong>rly to handle the resultant<br />

absolute (or dynamic) viscosity. Certain mines in Peru contain material called soft<br />

high clay, which can increase the absolute (or dynamic) viscosity of the <strong>slurry</strong> up to 400<br />

mPa at weight concentrations in excess of 45%. Dilution to lower concentration and<br />

changes to recycling load are solutions to such a problem.<br />

1-2-2 Heterogeneous Flows<br />

In a heterogeneous flow, solids are not uniformly mixed in the horizontal plane. A gradient<br />

of concentration exists in the vertical plane. Dunes or a sliding bed may form in<br />

the pi<strong>pe</strong>, with the heavier particles at the bottom and the lighter ones in sus<strong>pe</strong>nsion, particularly<br />

at the critical deposition velocity. The different phases retain their pro<strong>pe</strong>rties<br />

and the largest particles do not necessarily cause the biggest problems; it really de<strong>pe</strong>nds<br />

on the ratio that they are mixed with finer particles. Heterogeneous slurries are encountered<br />

in many placer mining, phosphate rock mining, and dredging applications. Concentration<br />

of particles remains low, typically less than 25% by weight in many dredging<br />

applications and below 35% by weight in many tailing disposal applications. Heterogeneous<br />

flows require a minimum carrier velocity. In some tailing applications of the<br />

Taconite mines of Minnesota, the typical deposition velocity is in excess of 3.4–4 m/s<br />

(11–13 ft/s).<br />

Nature being complex, flows are encountered that have the characteristics of heterogeneous<br />

or homogeneous flows. The concept of pseudohomogeneous flows is also used<br />

when a large fraction of particles are fine but there remain a sufficient fraction of coarse<br />

particles that may deposit as the flow s<strong>pe</strong>ed is reduced below a minimum value.<br />

1-2-3 Intermediate Flow Regimes<br />

Intermediate regimes occur when some of the particles are homogeneously distributed<br />

and others are heterogeneously distributed. Intermediate regime flows include tailings<br />

from mineral processing plants and a wide range of industrial slurries.<br />

1-2-4 Flows of Emulsions<br />

Strictly s<strong>pe</strong>aking, an emulsion is not <strong>slurry</strong>. An emulsion is a mixture of two phases at<br />

certain tem<strong>pe</strong>ratures resulting in an essentially homogeneous flow. An example of an<br />

emulsion is a mixture of bitumen at 70% by volume with water at 30% by volume. If sur-


factants are used, the bitumen remains well mixed with water in a certain tem<strong>pe</strong>rature<br />

range. Emulsions can become unstable under certain high shear rates or through very tight<br />

clearances of pumps according to Nunez et al. (1996). In the 1990s, PDVSA-BITOR constructed<br />

a 300 km (188 mi) pi<strong>pe</strong>line in Venezuela with a diameter of 660–915 mm (26–36<br />

in) for transporting their ORIMULSION fuel, a mixture of highly concentrated bitumen<br />

and water. The fuel is a substitute for coal in thermal plants.<br />

Emulsions do not encounter the deposition velocity of slurries, but as flows become<br />

unstable, fine droplets of heavy oils or bitumen may coalesce into larger ones, causing<br />

changes of flow.<br />

1-2-5 Flows of Emulsions—Slurry Mixtures<br />

A mixture of an emulsion and solids such as fine coal could be used to produce a fluid<br />

with a high calorific value. Coal must be very fine to burn readily in combustion furnaces.<br />

The flow is similar to a homogeneous flow with a high absolute (or dynamic) viscosity<br />

component.<br />

1-3 SINKING VELOCITY OF PARTICLES<br />

AND CRITICAL VELOCITY OF FLOW<br />

Various parameters of s<strong>pe</strong>ed determine whether a mixture may separate or continue to<br />

flow. In fact, the designer of a thickener or a mixer is often more interested in the sinking<br />

velocity of particles. On the other hand, the designer of a pi<strong>pe</strong>line has to pay attention to<br />

the critical velocity of flow, settling s<strong>pe</strong>ed, and whether the flow is vertical or horizontal,<br />

particularly in the case of heterogeneous flows.<br />

1-3-1 Sinking or Terminal Velocity of Particles<br />

This is the minimum s<strong>pe</strong>ed needed to maintain particles in sus<strong>pe</strong>nsion, particularly in a<br />

process of mixing or thickening. This velocity is not identical with the critical velocity of<br />

flow, and should not be confused with it.<br />

Table 1-8 presents examples of sinking velocity of various soils. The designer of a<br />

mixing system or a thickener is encouraged to conduct lab tests, since clays may be mixed<br />

with sands in some areas, or the soil may be stratified, with layers of different materials.<br />

1-3-2 Critical Velocity of Flows<br />

GENERAL CONCEPTS OF SLURRY FLOWS<br />

1.17<br />

In Chapters 3, 4, and 5 the mechanics of solid sus<strong>pe</strong>nsions are described in detail. An important<br />

parameter to introduce in this chapter is the critical velocity of a <strong>slurry</strong> flow. Figure<br />

1-8 plots the pressure loss <strong>pe</strong>r unit length on the y-axis, versus the velocity V of a <strong>slurry</strong><br />

flow on the x-axis. Five points are shown for flow at a constant volume concentration.<br />

For this <strong>slurry</strong> of moderate viscosity, the flow is stationary and the solids clog the pi<strong>pe</strong>line<br />

below point 1. There is insufficient s<strong>pe</strong>ed to move the particles. As the flow is accelerated,<br />

the s<strong>pe</strong>ed reaches point 1, which is called the deposition critical velocity V D, or minimum<br />

s<strong>pe</strong>ed to start the flow. Between points 1 and 2, the bed builds up, dunes form, and<br />

the different phases are well separated. Between points 2 and 3 the flow is streaking but


Pressure drop <strong>pe</strong>r unit of length<br />

1.18 CHAPTER ONE<br />

TABLE 1-8 Sinking Velocity of Soil Particles (after Sulzer Pumps, 1998, with<br />

<strong>pe</strong>rmission of Elsevier)<br />

Particle<br />

Soil grain size identification<br />

diameter, Mesh size, Sinking Sinking Grain size<br />

micrometers US fine velocity, m/s velocity, ft/s by ASTM Grain size, international<br />

0.2 3 × 10 –8 Clay Fine clay<br />

0.6 2.8 × 10 –7 Coarse clay<br />

1 7 × 10 –6<br />

2 9.2 × 10 –6<br />

5 17 × 10 –6 Silt Fine silt<br />

6 25 × 10 –6<br />

20 28 × 10 –5 Coarse silt<br />

50 270 17 × 10 –4 Fine sand Intermediate silt to sand<br />

60 230 25 × 10 –4<br />

100 150 0.07 Fine sand<br />

200 70 0.021 Medium coarse sand<br />

250 60 0.026 Coarse sand<br />

300 0.032<br />

500 35 0.053 Coarse sand<br />

600 30 0.063<br />

1000 18 0.10 Very coarse sand<br />

2000 10 0.17<br />

1<br />

<strong>slurry</strong><br />

2<br />

3<br />

water<br />

S<strong>pe</strong>ed of flow<br />

4<br />

5<br />

y/D<br />

concentration<br />

4 - 5<br />

Pseudohomogeneous<br />

3 - 4<br />

Jumping and rolling<br />

2 - 3<br />

Streaking<br />

1 - 2<br />

Bedding<br />

Below 1<br />

Stationary and<br />

clogging<br />

FIGURE 1-8 Pressure drop versus velocity for water and for a <strong>slurry</strong> mixture. (After Sulzer<br />

Pumps, 1998, with <strong>pe</strong>rmission of Elsevier.)


GENERAL CONCEPTS OF SLURRY FLOWS<br />

momentum is building up. Between points 3 and 4, there is sufficient s<strong>pe</strong>ed to cause<br />

jumping and rolling of the coarse particles. Above point 4, the s<strong>pe</strong>ed is sufficiently high<br />

to allow a pseudohomogeneous flow in which the fine particles act as a carrier for the<br />

coarse particles. These stages are extensively reviewed in Chapter 4.<br />

When absolute (or dynamic) viscosity is an important factor, such as in clayish slurries<br />

or homogeneous flows, another parameter, the viscous transition critical velocity V T must<br />

be determined. There are two regimes for flow of homogeneous mixtures. Flow at s<strong>pe</strong>eds<br />

less than V T is associated with laminar flows, whereas flows above V T are characteristic of<br />

turbulent flows.<br />

Flow in the laminar regime is often characterized by a friction loss factor, which is<br />

64/Re for the Darcy factor or 16/Re for the Fanning factor (this topic will be discussed in<br />

more details in Chapter 2). As a result, the losses in the laminar regime ap<strong>pe</strong>ar to be a linear<br />

function of s<strong>pe</strong>ed, whereas in the turbulent regime they are proportional to the square<br />

of the s<strong>pe</strong>ed. As we will see in Chapter 5, researchers have struggled with s<strong>pe</strong>cial definitions<br />

of a modified Reynolds number for non-Newtonian flows.<br />

Emulsions have been pum<strong>pe</strong>d over long distances in laminar flow. Nunez et al. (1996)<br />

demonstrates the existence of certain effects similar to comminution, i.e., breakup of<br />

large droplets into finer ones, as well as coalescence of small particles into larger ones under<br />

different flow regimes, shear rates, and constraints.<br />

A question often asked is what the relationship between the sinking s<strong>pe</strong>ed (as <strong>pe</strong>r<br />

Table 1-8) and the deposition of critical velocity V D? This question comes up when the<br />

rheology laboratory produces the results of thickening tests, and when there is not<br />

enough money or time to conduct pro<strong>pe</strong>r <strong>slurry</strong> loop tests. This point will be examined<br />

in Chapter 4 and various approaches have been adopted over the years, from the simplest<br />

assumption that the critical deposition velocity is 17 times as large as the terminal<br />

or sinking velocity, to more complex mathematical formulae. Often in a lab test, the<br />

coarse particles deposit rapidly while the fine particles are still in sus<strong>pe</strong>nsion. Pro<strong>pe</strong>r<br />

pump tests are often recommended, particularly when a multimillion dollar pi<strong>pe</strong>line is<br />

being designed. Tests can be conducted at a number of universities, provincial and state<br />

research centers, or with the help of manufacturers of <strong>slurry</strong> pumps. Examples include<br />

the Saskatchewan Science Research Center in Canada; the GIW research lab of KSB<br />

Pumps (USA), described by Wilson et al. (1992); the Slurry Research Lab of Mazdak<br />

International Inc. (U.S.A.); Texas A&M University (U.S.A.); and Melbourne University<br />

(Australia), among others.<br />

1-4 DENSITY OF A SLURRY MIXTURE<br />

The density of a <strong>slurry</strong> mixture is a function of<br />

� The density of the carrier fluid<br />

� The density of the solid particles<br />

� The concentration by volume of the solid phase<br />

1.19<br />

The density of the solid particles is determined carefully by various ex<strong>pe</strong>rimental<br />

methods. Fine particles tend to entrap air, which the lab technician must remove by pro<strong>pe</strong>r<br />

agitation or by adding a small quantity of wetting agent.<br />

Some materials exhibit a change of packing abilities and therefore density as a function<br />

of particle size. If the solids are to be passed through a comminution process, a SAG<br />

(semiautogeneous), or ball mill, they can occupy more volume <strong>pe</strong>r unit mass as they be-


1.20 CHAPTER ONE<br />

come finer. The <strong>slurry</strong> engineer is therefore encouraged to measure the solid’s density at<br />

the proposed size of particles to be transported in <strong>slurry</strong> form.<br />

Certain errors can occur in evaluating the density of solids with heterogeneous mixtures.<br />

If the heavier <strong>slurry</strong> particles settle out and a sample is taken, it may reflect a<br />

greater density of finer particles. Due to these possible sources of error, the engineer is<br />

encouraged to measure the density of the <strong>slurry</strong> mixture after pro<strong>pe</strong>r mixing, and to use<br />

the data on concentration by weight or by volume to work back to the density of the<br />

solids.<br />

The density of a <strong>slurry</strong> mixture is expressed as<br />

100<br />

�m = ���<br />

(1-4)<br />

Cw/�s + (100 – Cw)/�L where<br />

Cw = concentration by weight<br />

�m = density of the mixture phase<br />

�l = density of the liquid phase<br />

�s = density of the solid phase<br />

Engineers use the term concentration by weight, as it is easier to convert back into the<br />

total tonnage of solids to be transported through a pi<strong>pe</strong>line or across an extractive metallurgy<br />

plant. However, the characteristics of the mixture, the mechanics of flow, and the<br />

resultant physical pro<strong>pe</strong>rties are more related to the concentration by volume.<br />

The concentration by volume of solids in a mixture is expressed as<br />

C w� m<br />

Cv = � = ���<br />

(1-5)<br />

�s Cw/�s + (100 – Cw)/�L The concentration by weight of solids in a mixture is expressed as<br />

Cv�s �<br />

�m<br />

100 C w/� s<br />

Cv = = (1-6)<br />

Example 1-A<br />

A pi<strong>pe</strong>line is designed to transport 140 metric ton (308,000 lb) of sand <strong>pe</strong>r hour. The s<strong>pe</strong>cific<br />

gravity of the sand particles is 2.65 (or the density is 2650 kg/m3 ��<br />

Cv�s + (100 – Cv) ). The concentration<br />

by weight is 30%. Determine the density of the mixture if the carrier fluid is water and determine<br />

the resultant flow rate.<br />

� Weight of sand in the mixture over a <strong>pe</strong>riod of 1 hr is 140 metric ton (308,000 lb)<br />

� Weight of equivalent volume of water equivalent to the sand content is 140,000<br />

kg/2,650 kg/m 3 = 52,800<br />

� Weight of water in the <strong>slurry</strong> mixture at a concentration of 30% by weight = 140,000<br />

kg (100 – 30)/30 or 326,600 kg<br />

� Total weight of <strong>slurry</strong> mixture transported in 1 hr is the sum of the weight of water and<br />

sand or 140,000 + 326,600 = 466,600 kg/hr<br />

� Total weight of equivalent volume of water is 326,000 + 52,800 = 378,800 kg or 378<br />

m 3 of liquid, since density of water is 1000 kg/m 3<br />

The density of the <strong>slurry</strong> mixture is therefore 466,600 kg/378 m 3 = 1,230 kg/m 3 . Alternatively,<br />

the s<strong>pe</strong>cific gravity of the <strong>slurry</strong> compared to water is 1.23. The flow rate, being<br />

the volume <strong>pe</strong>r unit of time, is equivalent to 378 m 3 /hr.<br />

C v� s


1-5 DYNAMIC VISCOSITY OF A<br />

NEWTONIAN SLURRY MIXTURE<br />

Although density is essentially a static pro<strong>pe</strong>rty, absolute (or dynamic) viscosity is a dynamic<br />

pro<strong>pe</strong>rty and tends to reduce in magnitude as the shear rate in a pi<strong>pe</strong>line increases.<br />

Thus, engineers have had to define different forms of viscosity over the years, everything<br />

from dynamic viscosity, to kinematic viscosity, to effective pi<strong>pe</strong>line viscosity. The effective<br />

pi<strong>pe</strong>line viscosity will be discussed in detail in Chapters 3, 4, and 5. In this chapter,<br />

the reader is introduced to basic concepts of the mixture of <strong>slurry</strong> in a stationary state.<br />

This is effectively what the pump, or a mixer, might see at the start-up of a plant. As is often<br />

the case, when the driver cannot deliver enough torque to overcome the absolute (or<br />

dynamic) viscosity, the o<strong>pe</strong>rator is forced to dilute the <strong>slurry</strong> mixture.<br />

Plasticity as defined in Section 1-1-4 is an important parameter in determining overall<br />

absolute (or dynamic) viscosity of a mixture of clay and water. There are, however, numerous<br />

soils in nature, such as sand and water or gravel and water, in which the solids<br />

contribute little to the overall absolute (or dynamic) viscosity, except in terms of their<br />

concentration by volume.<br />

1-5-1 Absolute (or Dynamic) Viscosity of Mixtures with Volume<br />

Concentration Smaller Than 1%<br />

For such solid–liquid mixtures in diluted form, Einstein develo<strong>pe</strong>d the following formula<br />

for a linear relationship between absolute (or dynamic) viscosity and volume concentration:<br />

= 1 + 2.5� (1-7)<br />

where<br />

�m = absolute (or dynamic) viscosity of the <strong>slurry</strong> mixture<br />

�L = absolute (or dynamic) viscosity of the carrying liquid<br />

This is a very simple equation that is based on the following assumptions:<br />

� Particles are fairly rigid<br />

� The mixture is fairly dilute and there is no interaction between the particles<br />

Such a flow is not encountered, except in laminar regimes of very dilute concentrations<br />

(below a volume concentration of 1%).<br />

1-5-2 Absolute (or Dynamic) Viscosity of Mixtures with Solids<br />

with Volume Concentration Smaller than 20%<br />

Thomas (1965) took the equation of Einstein further by calculating for higher volumetric<br />

concentrations of Newtonian mixtures:<br />

� m<br />

� �L<br />

where K 1, K 2, K 3, and K 4 are constants<br />

GENERAL CONCEPTS OF SLURRY FLOWS<br />

� m<br />

� �L<br />

1.21<br />

= 1 + K 1� + K 2� 2 + K 3� 3 + K 4� 4 + . . . (1-8)


1.22 CHAPTER ONE<br />

K 1 is the Einstein constant of 2.5 (from Equation 1-7), and K 2 has been found to be in<br />

the range of 10.05–14.1 according to Guth and Simha (1936). It is difficult to extrapolate<br />

the higher terms K 3 and K 4 in Equation 1-8. They are ignored with volumetric concentrations<br />

smaller than 20%.<br />

1-5-3 Absolute (or Dynamic) Viscosity of Mixtures with High<br />

Volume Concentration of Solids<br />

For higher concentrations, Thomas (1965) proposed the following equation with an exponential<br />

function:<br />

= 1 + K 1� + K 2� 2 + A exp(B�) (1-9)<br />

where<br />

K2 = 10.05<br />

A = 0.00273<br />

B = 16.6<br />

Figure 1-9 is based on Equation 1-9 and is widely accepted in the <strong>slurry</strong> industry for heterogeneous<br />

mixtures of a Newtonian rheology.<br />

1-6 SPECIFIC HEAT<br />

Thomas (1960) derived an equation for the s<strong>pe</strong>cific heat of a mixture as a function of the<br />

s<strong>pe</strong>cific heat of the liquid and solid phases:<br />

1-7 THERMAL CONDUCTIVITY AND<br />

HEAT TRANSFER<br />

CpmCws + CpmCwL Cpm= ��<br />

(1-10)<br />

100<br />

Thermal conductivity is difficult to measure, as solids may settle during the test. Sometimes<br />

it is recommended to apply a small quantity of gel to maintain the solids in sus<strong>pe</strong>nsion.<br />

Orr and Dalla Valle (1954) derived the following equation for the thermal conductivity<br />

of <strong>slurry</strong> mixtures:<br />

where<br />

k = thermal conductivity<br />

and subscripts<br />

l = liquid<br />

m = mixture<br />

s = solids.<br />

� m<br />

� �L<br />

2kl + ks – 2�(kl – ks) ���<br />

2kl + ks – 2�(kl + ks) k m = k l� � (1-11)


1.23<br />

Ratio of viscosity of <strong>slurry</strong> mixture vs.<br />

L<br />

m<br />

viscosity of carrier liquid<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

CV<br />

Volumetric concentration of solids<br />

0 10 20 30 40 50 60 70 80 90<br />

C [%]<br />

V m L<br />

1 1.029<br />

3 1.089<br />

5 1.156<br />

7 1.233<br />

10 1.365<br />

12 1.465<br />

14 1.575<br />

16 1.696<br />

18 1.83<br />

20 1.978<br />

22 2.142<br />

25 2.426<br />

27 2.649<br />

29 2.907<br />

31 3.210<br />

33 3.573<br />

35 4.017<br />

37 4.570<br />

39 5.273<br />

42 6.734<br />

43 7.37<br />

44 8.103<br />

C [%]<br />

V<br />

m L<br />

45 8.950<br />

46 9.932<br />

47 11.07<br />

48 12.40<br />

49 13.94<br />

50 15.75<br />

51 17.86<br />

52 20.33<br />

53 23.22<br />

54 26.62<br />

55 30.61<br />

56 35.29<br />

57 40.80<br />

58 47.28<br />

59 54.91<br />

60 63.89<br />

61 74.47<br />

62 86.94<br />

63 101.63<br />

64 118.95<br />

65 139.4<br />

FIGURE 1-9 Ratio of viscosity of mixture versus viscosity of carrier in accordance with the Thomas equation for coarse slurries.


1.24 CHAPTER ONE<br />

Metzner et al. (1959, 1960) published articles on heat transfer for slurries. The applications<br />

for heat transfer problems have been confined to the nuclear industry, the processing<br />

of tar sands, feeding <strong>slurry</strong> to autoclaves for thermal processing, or certain emulsionbased<br />

slurries.<br />

1-8 SLURRY CIRCUITS IN EXTRACTIVE<br />

METALLURGY<br />

It would be beyond this book to discuss the principles of extractive metallurgy. Slurry is a<br />

very important component in the processing of ores to the final disposal of tailings and<br />

shipping of concentrate. Chapter 7 is dedicated to equipment for <strong>slurry</strong> processing. There<br />

are three main processes used for extractive metallurgy:<br />

1. Hydrometallurgy, which implies processing the ore using a liquid medium<br />

2. Electrometallurgy, which involves the application of electric and electro-chemical<br />

processes to extract the ore<br />

3. Pyrometallurgy, which involves the use of heat (roasting, smelting, etc.) for processing<br />

the ore<br />

The most common minerals of high metric tonnage (iron, aluminum, cop<strong>pe</strong>r, titanium,<br />

nickel, chromium, magnesium, zinc, etc.) are found in nature as oxides and sulfides and<br />

as a combination of both. Ores are sometimes a mixture of rich metal composition and<br />

poorer compositions called gangue. The gangue can be acidic or alkaline, and determines<br />

the ty<strong>pe</strong> of flux used for pyrometallurgy. Since ores come in all levels of complexity, various<br />

methods of processing have been develo<strong>pe</strong>d over the years.<br />

The first process ore undergoes is called classification or ore dressing. The purpose<br />

here is to separate the richer components of a mixture from the unuseful soils. Mineral<br />

processing may be used to produce a single stream, as is the case with taconite circuits,<br />

where iron extraction is the main activity. It may also create two streams, such as cop<strong>pe</strong>r<br />

concentrate and gold concentrate, when both minerals are found in the same ore<br />

body .<br />

Mineral processing is usually undertaken at the mine. Its purpose is to separate some<br />

of the gangue before shipment of a concentrate. The concentrate is richer in the desired<br />

mineral than the original soil.<br />

1-8-1 Crushing<br />

In Figure 1-10, a block diagram for crushing and grinding is presented. These are two of<br />

the fundamental steps taken to start a <strong>slurry</strong> circuit. Large rocks are first crushed to an acceptable<br />

size. De<strong>pe</strong>nding on the ty<strong>pe</strong> of equipment used, crushing may be done in a single<br />

step to feed a semiautogeneous mill, or in three steps (primary, secondary, and tertiary<br />

crushing).<br />

Rocks are transported from one crusher to another by conveying in a dry form. Their<br />

initial size of 300–600 mm (1–2 ft) is reduced to 100–150 mm (4–6 in). Jaw, gyrator, and<br />

cone crushers are commonly used during these stages. The crushed material is transported<br />

by conveyors to a storage area called the stockpile. From the stockpile, crushed rocks are<br />

transported to the grinding and milling circuit.


wells<br />

Reclaim Tailings<br />

Water Dam<br />

Lake<br />

River<br />

stockpile<br />

tank<br />

water<br />

30-40% solids<br />

Tailings pi<strong>pe</strong>s<br />

GENERAL CONCEPTS OF SLURRY FLOWS<br />

Ore


1.26 CHAPTER ONE<br />

which hel<strong>pe</strong>d eliminate complicated gearboxes for SAG mills. Some of the largest SAG<br />

mills are now built with a diameter of 12.2 m (40 ft).<br />

The process of crushing is essentially a dry process but the process of milling is a wet<br />

process in which <strong>slurry</strong> comes to play an important role. The use of water eliminates the<br />

dangerous generation of dust associated with dry grinding.<br />

It is not possible to undertake milling or wet grinding in a single step. The <strong>slurry</strong> is recirculated<br />

as follows. Initially, the coarse and fine particles are separated through a coarse<br />

screen or a s<strong>pe</strong>cial cyclone. Then the coarse particles are returned to the SAG mill and the<br />

intermediate sized particles are sent to the ball mill. The fine sized particles are taken<br />

from the cyclone overflow to a magnetic separation, electrostatic separation, or flotation<br />

plant. It is therefore not uncommon to recirculate 250–350% of the feedstock through the<br />

circuit of grinding and milling.<br />

Attention should be paid to the presence of clays in the ore, and associated dynamic<br />

viscosity. Flow from the ball mills can reach a concentration of 40–50% by weight and<br />

certain non-Newtonian rheology may manifest itself. Over the years, a plant that started<br />

in rocky ground may encounter more clay as it proceeds into dee<strong>pe</strong>r depths.<br />

1-8-3 Classification<br />

This is essentially a process to separate the particles according to their sinking rates in<br />

water. Wet classifiers are used with grinding and may include rubber-lined or ceramic cyclones<br />

(called hydrocyclones) and spiral mechanical classifiers.<br />

The principle of the cyclone is to feed the <strong>slurry</strong> tangentially and force it to rotate. By<br />

centrifugal forces, the coarser particles sink to the bottom of the cone while the finer particles<br />

float to the top. Both streams separate. The underflow, which consists of coarse material,<br />

and the overflow, which consists of fines, are then directed to other circuits. The<br />

underflow is fed back to the ball mills and the overflow is directed to the flotation circuit<br />

or other ty<strong>pe</strong>s of separators.<br />

The spiral mechanical classifier is used in pools. The heavier particles are allowed to<br />

settle in the pool while the finer particles float and flow out of the pool. The heavier particles<br />

are then removed from the bottom of the pool with a spiral or mechanical device.<br />

From cyclones or from the grinding circuit, the <strong>slurry</strong> may pass through different ty<strong>pe</strong>s<br />

of separators such as flotation circuits, electrostatic separators, and magnetic separators.<br />

Their purpose is to separate by chemical, electrical, or magnetic forces the minerals from<br />

the gangue. These steps occur before further thickening prior to feeding the pi<strong>pe</strong>line. The<br />

gangue is diverted to tailings circuits (see Figure 1-11).<br />

1-8-4 Concentration and Separation Circuits<br />

After a considerable effort to reduce the sizes of particles, it is important to separate the<br />

richer soils from the slimes or gangue. This step is achieved by using the pro<strong>pe</strong>rties of the<br />

ore itself.<br />

Gravity devices work on the principle that the ore (such as gold or diamonds) is heavier<br />

than the gangue. These devices include shaking units, the classic miner’s pan, rocking<br />

cradle devices, or more sophisticated gold concentrators or mineral sand concentrators.<br />

The drawback of these <strong>systems</strong> is that they may not necessarily be able to treat the fine<br />

particles produced by grinding and milling. In diamond extraction plants, X-ray machines<br />

are used in conjunction with gravity separation to detect the diamonds.<br />

Gravity devices are also used in a number of dry processes as well as <strong>slurry</strong> processes


1.27<br />

wells<br />

Lake<br />

River<br />

tank<br />

Reclaim<br />

Water<br />

Tailings<br />

Dam<br />

Tailings pi<strong>pe</strong>s<br />

30-40% solids<br />

Crushing, grinding, milling (see fig 1- 10)<br />

Flotation Electrostatic<br />

Cells Separators<br />

gangue<br />

tailings<br />

sump<br />

reclaim<br />

water<br />

Magnetic<br />

Separators<br />

minerals<br />

Gravity<br />

Separators<br />

tailings concentrate<br />

thickener<br />

thickener<br />

<strong>slurry</strong><br />

pi<strong>pe</strong>line<br />

filtering<br />

drying<br />

smelting<br />

burning<br />

concentrate<br />

at about 20%<br />

mineral<br />

FIGURE 1-11 Block fi 1diagram 11 for thickening and disposal of tailings, used as the basis for the design of <strong>slurry</strong> concentrate and<br />

tailings pi<strong>pe</strong>lines.


1.28 CHAPTER ONE<br />

for mineral sand and placer mining. In the case of mineral sands, the presence of a wide<br />

range of heavy oxides allows the miner to separate the various components by using gravity<br />

in conjunction with the magnetic or electrostatic pro<strong>pe</strong>rties.<br />

Magnetic devices are es<strong>pe</strong>cially useful, since iron is one of the most common minerals<br />

and it is possible to separate iron oxides from gangue by applying a magnet. This is usually<br />

done by introducing the <strong>slurry</strong> over a rotating drum, as shown in Figure 1-12; the drum<br />

picks up the magnetic concentrate while the remaining soils are diverted away by the flow<br />

of the remaining <strong>slurry</strong>. Magnetic devices are common in taconite processing plants, iron<br />

ore plants, as well as in mineral sand plants.<br />

Electrostatic devices were develo<strong>pe</strong>d in Australia to process beach and mineral sands.<br />

The sand ore is fed to a conducting and grounding rotor and is exposed to ionization. The<br />

particles, which have certain electrostatic pro<strong>pe</strong>rties, are attracted by the electric charge<br />

and are separated from the other particles. The nonconducting particles drop on the rotor<br />

and are brushed away into a separate container (Figure 1-13).<br />

Flotation devices use the principle of flotation to separate particles that are wettable<br />

from other particles. This is a very common process with sulfides but is less efficient with<br />

oxides. The cyclone overflow or the fine particles in the <strong>slurry</strong> after undergoing milling<br />

and grinding are fed to a series of flotation tanks (or a flotation machine). An agitator provides<br />

vigorous mixing. Air is introduced from a separate compressor line and chemical<br />

reagents are added to create froth. The nonwettable minerals float on top of the froth and<br />

are pum<strong>pe</strong>d away by froth-handling pumps or scra<strong>pe</strong>d away by mechanical devices. The<br />

wettable particles, such as the gangue, do not float on top of the froth and sink to the bottom<br />

of the tank. S<strong>pe</strong>cial tailing pumps may then pump away the gangue, sometimes for<br />

further grinding and processing (particularly gangue from the first flotation tank) and<br />

sometimes to the final tailing box (see Figure 1.14).<br />

The process of froth generation and flotation is more efficient when carried out in<br />

steps. A series of up to six tanks may be constructed to drop gradually, with launders in<br />

between.<br />

Without reagents, only graphite or molybdenum would be nonwettable. Reagents have<br />

been produced by the industry for different grades of sulfides, to depress or activate the<br />

extraction of certain minerals, to control pH, etc.<br />

Feed<br />

Other soils<br />

Magnetic<br />

concentrate<br />

rotating<br />

magnet<br />

FIGURE 1-12 A drum-ty<strong>pe</strong> magnetic separator. The drum is sometimes replaced by a magnetic<br />

belt on a s<strong>pe</strong>cial table.


GENERAL CONCEPTS OF SLURRY FLOWS<br />

Feed<br />

Nonconductor soils<br />

Nonionizing<br />

electrode<br />

fine-wire<br />

electrode<br />

ionizing<br />

Conductors<br />

concentrate<br />

FIGURE 1-13 An electrostatic separator.<br />

1.29<br />

The secondary grinding that is applied to the underflow of the circuits from flotation is<br />

useful for extracting secondary ores, and has been applied successfully in cop<strong>pe</strong>r–gold<br />

ores for further extraction of the gold.<br />

A mineral process plant includes gravity flows from hydrocyclones to ball mills and<br />

pum<strong>pe</strong>d flows from SAG mills to hydrocyclones. A good plant layout must allow space<br />

for repairs and long bends, and provide the ability to join and split flows. The use of<br />

three dimensional computer modeling is a very useful tool for the design engineer to determine<br />

the slo<strong>pe</strong> of launders and physical constraints to the layout of the plant (Figure<br />

1-15)<br />

froth bubbles with mineral concentrate<br />

FIGURE 1-14 A flotation circuit.<br />

aeration agitator<br />

pump<br />

secondary grinding


1.30 CHAPTER ONE<br />

1-8-5 Piping the Concentrate<br />

From these processes of classification, a concentrate is obtained. The <strong>slurry</strong> can be further<br />

thickened or dewatered using thickeners to a concentration of 50–60% by weight. The<br />

concentrate can then be pum<strong>pe</strong>d for hundreds of kilometers to a port from which it can be<br />

ship<strong>pe</strong>d to a pyrometallurgy plant. At the port, a filtering plant can provide further dewatering.<br />

Long pi<strong>pe</strong>lines are used to transport concentrate. At Cuajone, in Peru, an o<strong>pe</strong>n launder<br />

is used to transport tailings from an altitude of 3000 m (10,000 ft) down to sea level.<br />

The potential energy drop is used to overcome the friction losses of the launder. In<br />

Escondida, Chile, cop<strong>pe</strong>r concentrate flows by gravity from an altitude in excess of<br />

2500 m (8200 ft) above sea level over a distance in excess of 200 km (125 mi) to a port<br />

at sea level.<br />

Thus, in a typical cop<strong>pe</strong>r extraction process the rocks are reduced to very fine particles<br />

through vertimills, semiautogeneous mills, and ball mills. Further separation occurs<br />

through flotation circuits and grinding.<br />

1-8-6 Disposal of the Tailings<br />

Once the concentrated ore has been extracted, the plant is left with the sands and slimes.<br />

These are dewatered to an acceptable concentration by weight of 35–45%. Using thicken-<br />

FIGURE 1-15 Three-dimensional computer representation of a grinding circuit with one<br />

central SAG mill, a ball mill on each side, hydrocyclones at the top left, and pumps (at the floor<br />

level). Courtesy of Hatch & Associates, Vancouver, Canada.


GENERAL CONCEPTS OF SLURRY FLOWS<br />

ers, the flow separates into clarified water that is returned to the plant for use in the<br />

milling and grinding circuits, and into an underflow of concentrated tailings. The underflow<br />

from the tailings is then pum<strong>pe</strong>d away to a large disposal pond, called a tailings dam.<br />

The sand sinks to the bottom of the tailings pond and the water from the top is returned<br />

back to the plant for further use.<br />

During the process of disposing of the tailings, spigots and other devices are used to<br />

separate coarse from fine particles at the discharge point; the coarse particles are used to<br />

build the beaches or the wall of the dam. Dams have been built up to a height of 200 m<br />

(656 ft). The environmental engineer must make sure that no dangerous chemicals seep<br />

through the ground. If the tailings contain dangerous substances, a plastic or clay (which<br />

tends to create a seal) lining for the pond may be recommended to prevent seepage. Some<br />

dangerous collapses of tailing dams have been reported over the years, with detrimental<br />

consequences when they contained cyanide products, as in the case of certain gold mines.<br />

The technology used in tailing pumps has improved. The casings can be designed to<br />

withstand 6.9 MPa (1000 psi) of pressure. However, these are essentially single-stage<br />

centrifugal pumps installed in series. Up to seven pumps have been installed in series in<br />

mines such as National Steel in Minnesota and Kelian Gold in Indonesia.<br />

1-9 CLOSED AND OPEN CHANNEL FLOWS,<br />

PIPELINES VERSUS LAUNDERS<br />

1.31<br />

The reader will find that over the years the majority of references on <strong>slurry</strong> flows have focused<br />

on pi<strong>pe</strong>lines because the interest in the field has concentrated on the ability to haul<br />

coal, sand, and phosphate hydraulically.<br />

Many mines, particularly in Chile and Peru, are located at very high altitudes. This demand<br />

has increased interest in gravity flows. In the early 1970s, Southern Peru Cop<strong>pe</strong>r installed<br />

one of the first long, o<strong>pe</strong>n launders to dispose of tailings to the sea. The launder<br />

was of a concrete and fiberglass design, with a U-sha<strong>pe</strong>d cross section. Another example<br />

of a long gravity pi<strong>pe</strong>line for cop<strong>pe</strong>r concentrate is the Escondida concentrate pi<strong>pe</strong>line in<br />

Chile, which is longer than 200 km (125 mi).<br />

Despite the increasing importance of long gravity pi<strong>pe</strong>lines, equipment has not kept up<br />

with the expanding need. Cave (1980) described tests on <strong>slurry</strong> turbines. A 350 mm × 300<br />

mm (14� × 12�) was reported by Burgess and Abulnaga (1991).<br />

Launders play a very important role in <strong>slurry</strong> flows of plants. Cyclone underflow is directed<br />

to ball mills then to SAG mills by gravity. Flows in these circuits can cause<br />

tremendous wear if provision is not made to control s<strong>pe</strong>eds. Launders in plants are typically<br />

rubber lined. Long-distance pi<strong>pe</strong>lines are manufactured of rubber-lined steel or extra-thick,<br />

high-density polyethylene (HDPE). Because of the importance of o<strong>pe</strong>n launders,<br />

gravity flows, and drop boxes, Chapter 6 is dedicated to these complex flows.<br />

Obviously, not all mines are located on mountaintops, and <strong>slurry</strong> pi<strong>pe</strong>line flow will<br />

continue to be the main emphasis of researchers. In long pi<strong>pe</strong>lines, centrifugal pumps can<br />

be installed at regular intervals; these require power to be brought in. For long-distance<br />

pumping, positive displacement pumps com<strong>pe</strong>te well with centrifugal pumps. The positive<br />

displacement pumps are of a diaphragm or hose design. They are extremely ex<strong>pe</strong>nsive.<br />

A 17.3 MPa (2500 psi) pump may range in price between U.S. $600,000 and<br />

$1,200,000 in year 2000 dollars. The higher capital investment required for positive displacement<br />

pumps is offset by their higher efficiency. These pumps are built to much<br />

smaller flow capacity than are large centrifugal pumps.


1.32 CHAPTER ONE<br />

1-10 HISTORICAL DEVELOPMENT OF<br />

SLURRY PIPELINES<br />

One of the first large engineering projects that involved transportation of solids by liquid<br />

was the dredging for the Suez Canal in the 1860s in Egypt. It was reported to have used<br />

conduits to dispose of the sand–water mixture. Nora Blatch in 1906 was probably the first<br />

<strong>pe</strong>rson to conduct a systematic investigation of the flow of solid–water mixtures. She<br />

used a 25 mm (1 in) horizontal pi<strong>pe</strong> and measured the pressure gradients as a function of<br />

flow, density, and solid concentration. As a result, between 1918 and 1924, a 200 mm (8<br />

in) pi<strong>pe</strong>line was installed in the Hammersmith power station, in London, England to<br />

transport coal <strong>slurry</strong> over a distance of 660 yd.<br />

In 1948, in France, the Institute of Research SOGREAH began a series of tests on<br />

transporting sand and gravel in pi<strong>pe</strong>s with a diameter from 38–250 mm (1.5–10 in). These<br />

extensive tests were the basis for the formulation by Durand of a number of equations that<br />

will be reviewed in Chapter 4. These equations have been subject to further refinements<br />

over the last 50 years.<br />

In 1952, in the United Kingdom, the British Hydromechanic Research Association<br />

(BHRA) started to study the hydraulic transport of lump coal, sand, gravel, and limestone.<br />

Limestone pi<strong>pe</strong>lines were constructed in Trinidad and England in 1960. The Trinidad<br />

pi<strong>pe</strong>line had a diameter of 204 mm (8 in), a length of 9.6 km (6 mi), and was designed to<br />

o<strong>pe</strong>rate in a laminar flow regime. The limestone pi<strong>pe</strong>line in England had a diameter of<br />

250 mm (10 in) and was 112 km (70 mi) long.<br />

In 1950, the Consolidated Coal Co. in the United States started to conduct research on<br />

the hydrotransport of fine “nonsettling” slurries. Concentrated coal with a weight concentration<br />

of 60% and particle size between minus 1168 �m (14 mesh) and minus 43 �m<br />

(325 mesh) was transported. The pi<strong>pe</strong>line transported 1.5 million tons of coal each year<br />

between 1957 and 1964. The pi<strong>pe</strong>line stretched 176 km (110 mi) from Cadiz, Ohio to<br />

Eastlake in Cleveland, Ohio.<br />

In 1957, the Colorado School of Mines collaborated with the American Gilsonite<br />

Company and designed a pi<strong>pe</strong>line with a diameter of 200 mm (8 in) to transport crushed<br />

gilsonite. The pi<strong>pe</strong>line was constructed between Bonanza, Utah and Grand Junction, Colorado.<br />

The particle size was minus 4.7 mm (4 mesh) and solids were pum<strong>pe</strong>d at a weight<br />

concentration of 48%. Two other pi<strong>pe</strong>lines were built in Georgia to transport kaolin in the<br />

1960s.<br />

In 1967, an iron ore concentrate <strong>slurry</strong> pi<strong>pe</strong>line started to o<strong>pe</strong>rate in Tasmania,<br />

Australia. The pi<strong>pe</strong>line had a diameter of 245 mm (9 5 – 8 in). Concentrate was transported<br />

at a weight concentration of 60% with an average particle size of minus 149 �m (100<br />

mesh) over a distance of 85 km (53 mi) through extremely rugged terrain (see Figure<br />

1-16).<br />

In 1970, the Black Mesa Pi<strong>pe</strong>line, one of the longest pi<strong>pe</strong>lines ever built up to that<br />

time, started o<strong>pe</strong>ration between the Black Mesa Coal fields in Arizona and the Mohave<br />

Power Plant in Nevada. Coal was ground to a particle size of minus 1168 �m (14 mesh),<br />

and transported in a pi<strong>pe</strong> with a diameter of 457 mm (18 in) over a distance of 437 km<br />

(273 mi). Coal was dewatered at the end of the line through a mill before combustion with<br />

preheated air.<br />

Since the 1970s, a number of short and long <strong>slurry</strong> pi<strong>pe</strong>lines have been constructed.<br />

Table 1-9 lists a number of such achievements. Now, at the beginning of the 21st century,<br />

new complex, multiphase tar–sand pi<strong>pe</strong>lines are planned for northern Alberta,<br />

Canada.


GENERAL CONCEPTS OF SLURRY FLOWS<br />

1.11 SEDIMENTATION OF DAMS—<br />

THE ROLE OF THE SLURRY ENGINEER<br />

1.33<br />

FIGURE 1-16 Long <strong>slurry</strong> pi<strong>pe</strong>lines must travel through isolated areas over long distances<br />

and may involve pressures up to 2500 psi that require s<strong>pe</strong>cial positive displacement pumps.<br />

Courtesy of Wirth Pumps, Germany.<br />

In the last 150 years, world population has grown fast and our modern standards of living<br />

de<strong>pe</strong>nd on the production of electricity, and production of food for at least two seasons a<br />

year. In an effort to meet these demands, engineers have built small as well as very large<br />

dams. In certain areas, very large man-made lakes have been dug in the earth, such as behind<br />

the Aswan High Dam in Egypt, the Ataturk Dam in Turkey, and new dams on the<br />

Yellow River in China.<br />

Some large rivers transport silt that tends to separate from water when the s<strong>pe</strong>ed of the<br />

flow is interrupted by a dam. This phenomenon is called siltation of dams. The problem


1.34 CHAPTER ONE<br />

TABLE 1-9 Examples of Slurry Pi<strong>pe</strong>lines Built Since 1957<br />

Site of pi<strong>pe</strong>line,<br />

Pi<strong>pe</strong><br />

diameter<br />

Pi<strong>pe</strong>line length<br />

Solids<br />

transported,<br />

million short Start-up<br />

Ore or name of pi<strong>pe</strong>line inch Mile km tons/yr date<br />

Coal Consolidation, USA 10 108 175 1.3 1957<br />

Black Mesa 18 273 440 4.8 1970<br />

ETSI 38 1036 1675 25 1979<br />

ALTON 24 180 112 10 1981<br />

Belonovo–Novosibirsk, Siberia,<br />

Russia<br />

158 256 3.4 1985<br />

Iron Savage River 9 53 86 2.25 1967<br />

concentrate Waipipi (Iron Sands) 8,12 6 9.7 1.0 1971<br />

Pena, Colorado 8 28 45 1.8 1974<br />

Las Truchas, Mexico 10 16 27 1.5 1976<br />

Sierra Grande, Argentina 8 20 32 2.1 1976<br />

Samarco, Brazil 20 244 395 12 1977<br />

Chongin, North Korea ? 61 98 4.5 1975<br />

New Zealand Sands, NZ 12 5 8<br />

Jian Shan, China 10 62 100<br />

La Parla–Hercules, Mexico 8/14 52/182 85/295 4.5 1982<br />

Cop<strong>pe</strong>r ore Los Bronces 24 35 56<br />

Cop<strong>pe</strong>r Bougainville, PNG 6 20 32 1.0 1972<br />

concentrate West Irian, Indonesia 4 69 111 0.3 1972<br />

Pinto Valley 4 11 17 0.4 1974<br />

OK Tedi, Papua New Guinea 6 96 155 1987<br />

Escondida, Chile (gravity line) 9 102 165 ? 1994<br />

Collahausi, Chile 7 125 203 1.0 1999<br />

Freeport, Indonesia 5 71 115<br />

Batu Hijau, Indonesia 6 11 18 1999<br />

Alumbrera, Argentina 6 194 314 0.8 1998<br />

Cop<strong>pe</strong>r Bougainville, Papua NG 34 31 50<br />

tailings Southern Peru Cop<strong>pe</strong>r (gravity) 150 1972<br />

Limestone Rugby 10 57 92 1.7 1964<br />

Calaveras 7 17 27 1.5 1971<br />

Michigan Limestone Tailings 20 1.2 2<br />

Phosphate Chevron, Vernal, Wyoming 10 94 152 1.3–2.5 1986<br />

ore concentrate Simplot 8 89 145<br />

Wenglu 8 28 45<br />

Dredging Dallas White Rock Lake, USA 24 33 21 11,000 gpm 1998


GENERAL CONCEPTS OF SLURRY FLOWS<br />

1.35<br />

of siltation of dams has not been well documented or studied. Chanson (1998) and Chanson<br />

and James (1999) examined the siltation of Australian dams. Certain dams in Australia<br />

became gradually fully silted between 1890 and 1960. They reported that the<br />

Koorawatha dam in New South Wales (Figure 1-17) became fully silted with bed-load<br />

material. The Cunningham Creek Dam in New South Wales, Australia was well studied<br />

by Hellstrom (1941) according to Chanson. Sedimentation problems are more acute with<br />

small dams than with medium-size and large reservoirs (Chanson and James,1999). Siltation<br />

at Eildon in the State of Victoria occurred in 1940 after torrential rainfalls following<br />

bushfires that had destroyed 50% of the catchment forest. The siltation at Eppalock in<br />

Victoria followed extensive gold mining, tree clearing, and hydraulic mining between<br />

1851 and 1890.<br />

There were some extreme siltation cases. The Quipolly Reservoir No. 1, in Australia<br />

underwent very rapid sedimentation between 1941 and 1943 at a rate in excess of<br />

1143m 3 /km 2 /year (9600 ft 3 /mi 2 /y). The Korrumbyn Creek Dam sedimented in less than 7<br />

years.<br />

Since the 1950s, improvements in land management practices and a better understanding<br />

of the problems of soil erosion have resulted in better approaches to the protection of<br />

dams.<br />

FIGURE 1-17 The Koorawatha Dam in Australia—fully silted. [From Chanson (1999).<br />

Reprinted by <strong>pe</strong>rmission of Butterworth-Heinemann.]


1.36 CHAPTER ONE<br />

Chanson and James (1998) discussed the hazards of fully silted dams. The weight of<br />

silt pressing against the concrete structure introduces a medium with a s<strong>pe</strong>cific gravity<br />

larger than the water for which the dam was designed. It is important in these cases to<br />

monitor the structure.<br />

In the 1970s and 1980s, the drought in Ethiopia and the Sudan reduced the flow of the<br />

waters to the Nile to dramatically low levels. Egypt avoided famine by using its massive<br />

man-made lake behind the Aswan High Dam (Lake Nasser). On the other hand, any s<strong>pe</strong>cialist<br />

who visits Egypt can feel that the fellahin (farmers) are s<strong>pe</strong>aking with nostalgia of<br />

the “Tamye” or silt that used to come with the annual flood and enrich the land. This raw<br />

material was the basis of natural nutrients, as well as mud for the construction of houses<br />

and manufacture of bricks.<br />

The Egyptian case is not unique. Certain dams in the United States are now the subject<br />

of discussions on decommissioning and some were removed in the 1990s. The <strong>slurry</strong> engineer<br />

can offer some much needed solutions. The twenty-first century will see <strong>slurry</strong> engineers<br />

providing adequate solutions in terms of dredging the lakes that are sedimenting<br />

and transporting the dredged silt to traditional lands, or to arid lands via s<strong>pe</strong>cial <strong>slurry</strong><br />

pi<strong>pe</strong>lines. A simple concept for such a solution is presented in Figure 1-18. It is proposed<br />

that in certain areas, particularly where the accumulation of silt is likely to apply pressure<br />

on the dam structure, submersible <strong>slurry</strong> pumps be installed on a <strong>pe</strong>rmanent basis. Dredging<br />

boats with dredging arms or submersible pumps and cutters would be used on the rest<br />

of the lake. Where the capital investment does not justify it, small dredgers with submersible<br />

pumps should be used. The <strong>slurry</strong> from these o<strong>pe</strong>rations will be dilute, it would<br />

be pum<strong>pe</strong>d to the shore through a floating plastic pi<strong>pe</strong>. It may be pum<strong>pe</strong>d in pi<strong>pe</strong>lines and<br />

diverted to canals for agricultural purposes. It may also be pum<strong>pe</strong>d to brick plants, where<br />

it would be dewatered and the silt used as a raw material.<br />

dredger<br />

Agriculture<br />

Submersible<br />

pump<br />

Dam<br />

Brick<br />

manufacture<br />

Fi 118<br />

FIGURE 1-18 Simplified flow sheet to remove silt behind dams. Silt would be dredged using<br />

submersible pumps at predetermined locations or in association with boat dredgers. The silt<br />

would be pum<strong>pe</strong>d in <strong>slurry</strong> form to agricultural farmlands or to s<strong>pe</strong>cial plants for the manufacture<br />

of bricks.<br />

silt


Slurry engineers can provide economic solutions to the siltation of dams. This effort<br />

should be made in conjunction with environmental engineers, as new eco<strong>systems</strong> often<br />

form around large dams. The author ho<strong>pe</strong>s that this <strong>handbook</strong> will be useful to the decision<br />

makers who have to deal with siltation of dams, while satisfying the concerns of environmental<br />

engineers, as decommissioning is not always the solution.<br />

1-12 CONCLUSION<br />

In this first chapter, some of the basic pro<strong>pe</strong>rties of solids, which are important to the<br />

composition of slurries, were reviewed. Their importance will be emphasized in the next<br />

few chapters. They may lead to Newtonian as well as more complex non-Newtonian<br />

flows that require s<strong>pe</strong>cial equations to determine the friction factors, velocity of flow,<br />

pi<strong>pe</strong> sizes, head, and efficiency losses in pumps.<br />

Wear is an important cost to be paid for transporting solids by liquids. This will be discussed<br />

later in the book when exploring <strong>slurry</strong> pumps and pi<strong>pe</strong>lines. The modern <strong>slurry</strong><br />

engineer can serve the mining and power industries by making possible the transportation<br />

of minerals, coal, coal–crude oil mixtures over very long distances, and also play a major<br />

role in dredging sediments behind dams to avoid dam failure and to provide arid lands<br />

with much needed silt.<br />

1-13 NOMENCLATURE<br />

GENERAL CONCEPTS OF SLURRY FLOWS<br />

A Constant<br />

B Constant<br />

Cc Coefficient of curvature<br />

Cu Coefficient of uniformity<br />

Cv Concentration by volume of the solid particles in <strong>pe</strong>rcent<br />

Cw Concentration by weight of the solid particles in <strong>pe</strong>rcent<br />

Cp Heat capacity<br />

d10 Grain size at which 10% of the soil is finer<br />

d30 Grain size at which 30% of the soil is finer<br />

d50 Grain size at which 50% of the soil is finer<br />

d60 Grain size at which 60% of the soil is finer<br />

d80 Grain size at which 80% of the soil is finer<br />

K1, K2, K3, K4 polynomial coefficients in Einstein’s equation for dynamic viscosity<br />

k Thermal conductivity<br />

LL Liquid limit of clay and silt soils<br />

PI Plastic index of clay and silt soils<br />

PL Plastic limit of clay and silt soils<br />

Re Reynolds number<br />

VD Deposition critical velocity<br />

VT Viscous transition critical velocity<br />

� Concentration by volume in decimal points<br />

� Absolute (or dynamic) viscosity<br />

� Density<br />

1.37


1.38 CHAPTER ONE<br />

Subscripts<br />

l Liquid<br />

m Mixture<br />

s Solids<br />

1-14 REFERENCES<br />

The American Society for Testing of Materials. 1993. Practice for Description and Identification of<br />

Soils (Visual–Manual Aggregate Mixtures). Standard D2488. Philadelphia: The American Society<br />

for Testing of Materials.<br />

The American Society for Testing of Materials. 1983. Test Method for Lime Content of Uncured<br />

Soil–Lime Mixtures. Standard D3155. Philadelphia: The American Society for Testing of Materials.<br />

The British Standard Institute. 1999. Code for Practice of Site Investigation. Standard BS 5930. London:<br />

The British Standard Institute.<br />

The British Standard Institute. 1999. Aggregate Abrasion Value. Standard BS 812. Pt 113. London:<br />

The British Standard Institute.<br />

Broch, E., and J. A. Franklin. 1972. The Point-Load Strength Test. International Journal for Rock<br />

Mechanics and Mineral Sciences, 9, 669–697.<br />

Burgess K. E, and B. E. Abulnaga.1991.The Application of Finite Element Methods to Warman<br />

Pumps and Process Equipment. Pa<strong>pe</strong>r presented to the Fifth International Conference on Finite<br />

Element Analysis in Australia, University of Sydney, Australia (July).<br />

Cave I. 1980. Slurry Turbines for Energy Recovery. In Seventh International Conference on the Hydraulic<br />

Transport of Solids in Pi<strong>pe</strong>lines, Sendai, Japan, pp. 9–15, Cranfield, United Kingdom:<br />

BHRA Group.<br />

Chanson H. 1998. Extreme Reservoir Sedimentation in Australia: A Review. International Journal<br />

of Sediment Research, UNESCO-IRTCES, 13, 3, 55–63.<br />

Chanson H. 1999. The Hydraulics of O<strong>pe</strong>n Channel Flows—An Introduction. Oxford, UK: Butterworth-Heinemann.<br />

Chanson H., and D. P. James.1999. Siltation in Australian Reservoirs: Some Observations and Dam<br />

Safety Implications.” In Proceedings 28th IAHR Congress, Graz, Austria, Pa<strong>pe</strong>r B5.<br />

Guth, E., and A. R. Simha. 1936. Viscosity of sus<strong>pe</strong>nsions and solutions. Kolloid-Z, 74, 266. Quoted<br />

in Wasp, E. J., J. P. Kenny, and R. L. Gandhi. 1977. Solid–Liquid Flow—Slurry Pi<strong>pe</strong>line<br />

Transportation. Aedermannsdorf, Switzerland: Trans. Tech. Publications.<br />

Hellstrom B. (1941) Nagra Lakttagelser over Vittring Erosion och Slambidning i Malaya och Australien.”<br />

Geografiska Annaler (Stockholm, Sweden), Nos. 1–2, pp. 102–124 (in Swedish).<br />

Herbrich, J. 1991. Handbook of Dredging Engineering. New York: McGraw-Hill Inc.<br />

The International Journal for Rock Mechanics and Mineral Sciences, 1979, 16, 141–156.<br />

The International Society for Rock Mechanics. 1977. Suggested Methods for Determining the<br />

Strength of Rock Materials in Triaxial Compression. Lisbon, Portugal: The International Society<br />

for Rock Mechanics.<br />

The International Society for Rock Mechanics. 1978. Suggested Methods for Determining the Deformability<br />

of Rock. Lisbon, Portugal: The International Society for Rock Mechanics.<br />

Metzner, A. B., and P. S. Friend. 1959. Heat Transfer to Turbulent Non-Newtonian Fluids. Ind. &<br />

Eng. Chem., 51, 7 (July), 879–882.<br />

Metzner, A. B., and D. F. Gluck. 1960. Heat Transfer to Non-Newtonian Fluids Under Laminar Flow<br />

Conditions. Chem. Eng. Science, 12, 3 (June), 185–190.<br />

Nunez, G. A., M. Briceno, C. Mata, and H. Rivas. 1996. Flow Characteristics of Concentrated Emulsions<br />

of Very Viscous Oil in Water. Journal of Rheology, 40, 3 (May/June), 405–423.<br />

Orr, C., and J. M. Dalla Valle. 1954. Heat-Transfer Pro<strong>pe</strong>rties of Liquid–Solid Sus<strong>pe</strong>nsions. Chem.<br />

Eng. Prog., Symp. Series No. 9, 50, 29–45.<br />

The Permanent International Association of Navigation Congresses. 1972. Classification of Soils to<br />

be Dredged. In Bulletin No. 11, Vol. I. The Permanent International Association of Navigational<br />

Congresses.


GENERAL CONCEPTS OF SLURRY FLOWS<br />

1.39<br />

Sargent, J. H. 1984. Classification of Soils to be Dredged. In Supplement to Bulletin No 47. The Permanent<br />

International Association of Navigation Congresses.<br />

Sulzer Pumps. 1998. Centrifugal Pump Handbook. New York: Elsevier.<br />

Thomas, D. G. 1960. Heat and Momentum Transport Characteristics of Non-Newtonian Aqueous<br />

Thorium Oxide Sus<strong>pe</strong>nsions. AIChE Journal, 6 (December), 631–639.<br />

Thomas D. G. 1965. Transient Characteristics of Sus<strong>pe</strong>nsions: Part VIII. A Note on the Viscosity of<br />

Newtonian Sus<strong>pe</strong>nsions of Uniform Spherical Particles. Journal Colloid Science, 20, 267.<br />

Wasp, E. J., J. P. Kenny, and R. L. Gandhi. 1977. Solid–Liquid Flow—Slurry Pi<strong>pe</strong>line Transportation.<br />

Aedermannsdorf, Switzerland: Trans Tech Publications.<br />

Wilson, K. C., G. R. Addie, and R. Clift. 1992. Slurry Transport Using Centrifugal Pumps. New<br />

York: Elsevier Applied Sciences.<br />

Further Reading:<br />

Wilson, G. 1976. Construction of Solids-Handling Centrifugal Pumps. In Pump Handbook. Edited<br />

by J. Karassik et al. New York: McGraw-Hill.


CHAPTER 12<br />

FEASIBILITY STUDY<br />

FOR A SLURRY PIPELINE<br />

AND TAILINGS<br />

DISPOSAL SYSTEM<br />

12-0 INTRODUCTION<br />

A consultant engineer has to convince his clients of the merits of a <strong>slurry</strong> pi<strong>pe</strong>line over alternative<br />

methods of transportation, whether it is for tailings disposal or concentrate shipping.<br />

One very important step in the design of a <strong>slurry</strong> pi<strong>pe</strong>line is to appreciate the economics<br />

involved in the process. There is no question that this is the effort of a team of<br />

engineers, geologists, and accountants.<br />

It is therefore very important to appreciate the different and complex facets of a feasibility<br />

study. The exercise of a feasibility study or basic engineering should go through a<br />

number of steps, or follow a kind of checklist. In this chapter, the different steps are presented<br />

for this purpose. A pi<strong>pe</strong>line for the disposal of tailings may be a few kilometers or<br />

miles long, whereas a <strong>slurry</strong> concentrate pi<strong>pe</strong>line may be few hundred kilometers or miles<br />

long. The role of the geologist or foundation engineer is critical to the successful construction<br />

of a tailings dam. It would be beyond the sco<strong>pe</strong> of this book to discuss geophysics.<br />

In recent years, there has been a trend toward disposing of tailings in the sea. Whether<br />

tailings are disposed over land or in the sea, there are environmental concerns that must<br />

be satisfied. The engineer should be aware of these issues. The presence of some corrosion<br />

inhibitors, cyanide, or toxic materials in the tailings must be handled carefully. Tailings<br />

dams are sometimes within reach of agricultural fields and seepage could have negative<br />

effects on the quality of underground water. Environmental concerns may represent<br />

hidden costs with particular re<strong>pe</strong>rcussions on <strong>slurry</strong> projects.<br />

This chapter presents an overview of basic engineering for a feasibility study. The<br />

study consists of identifying the components of a pi<strong>pe</strong>line (such as feeding station, main<br />

and booster stations, emergency dump ponds, final disposal tailings pond, and area for<br />

sub-sea disposal), the size of the pi<strong>pe</strong>line based on the anticipated flow rate, and the material<br />

of the pi<strong>pe</strong> based on pressure, chemical attack, erosion, and corrosion. Other as<strong>pe</strong>cts<br />

of the study outside the sco<strong>pe</strong> of the <strong>slurry</strong> engineer and which cannot be covered in this<br />

book involve the geological survey, the cost of excavation, the cost of construction, power<br />

lines, power stations, transformer stations, and SCADA or control <strong>systems</strong>. The s<strong>pe</strong>cialists<br />

involved in these areas rely on the <strong>slurry</strong> engineer for extensive information and<br />

12.1


12.2 CHAPTER TWELVE<br />

help in basic engineering by providing data on stability of soil, difficulty of the terrain,<br />

cost of power transmission, etc. In turn, this collaborative information is fed to the estimators<br />

and the project managers. The <strong>slurry</strong> engineer will be requested to make suggestions,<br />

review the feasibility study, and help purchase the equipment.<br />

12-1 PROJECT DEFINITION<br />

At an early stage of the feasibility study, the project is defined in the following terms:<br />

� Volume of <strong>slurry</strong> to be transported over the life of the project<br />

� Annual pum<strong>pe</strong>d flow of <strong>slurry</strong><br />

� Starting point of the pi<strong>pe</strong>line, such as a smelter or tailings dam, and a final point of the<br />

pi<strong>pe</strong>line, such as a port for export of the concentrate or power plant for burning coal<br />

� Proposed contour of the pi<strong>pe</strong>line<br />

� Existing roads and need for new roads for access to the pi<strong>pe</strong>line, tailings dam, or<br />

booster station<br />

� Proposed pressure rating of the pi<strong>pe</strong>line and the number of main and booster stations<br />

� Rheology of the <strong>slurry</strong><br />

� Environmental impact of the project<br />

� Stability of soil along the contour of the pi<strong>pe</strong>line and possibility of seismic problems or<br />

landslides<br />

� Need for a dewatering plant at the end of the pi<strong>pe</strong>line<br />

� Need for local generation of electricity for booster pump stations or for reclaim water<br />

stations<br />

� Need for a reclaim water pi<strong>pe</strong>line to return water to the starting point of the <strong>slurry</strong><br />

pi<strong>pe</strong>line<br />

� Estimation of excavation costs if the pi<strong>pe</strong>line is buried or if electric conduits are underground<br />

� Estimation of costs for power poles to transmit electricity<br />

� Protection of the pi<strong>pe</strong>lines from freezing in cold environments<br />

� Allowance for water hammer and transients<br />

� Allowance for thermal expansion in hot climates<br />

� Required modifications to existing thickeners, or filtering and dewatering plants as<br />

part of expansions of production and pum<strong>pe</strong>d flow rates<br />

� Mitigation against erosion, abrasion, and corrosion<br />

� Required purchase of land for pi<strong>pe</strong>line contour, tailings dam, dewatering plant, and<br />

booster stations<br />

� Engineering costs<br />

� Construction costs<br />

A general schematic diagram for the tailings disposal pi<strong>pe</strong>line (Figure 12-1) or the concentrate<br />

pi<strong>pe</strong>line (Figure 12-2) is made at an early stage to define the major components<br />

of the pi<strong>pe</strong>line.


Mineral<br />

Process<br />

Plant<br />

<strong>slurry</strong><br />

thickener<br />

Emergency<br />

pond<br />

isolation<br />

valve<br />

clarified<br />

water<br />

dilution<br />

water<br />

tailings<br />

pi<strong>pe</strong>line<br />

feed<br />

sump<br />

reclaim water pi<strong>pe</strong>line<br />

pump station<br />

(usually barge mounted)<br />

corrosion<br />

inhibitors<br />

pi<strong>pe</strong>line feed pump station<br />

(from 1 to 9 pumps in series<br />

up to 7.7 MPa (1100 psi))<br />

tailings<br />

pi<strong>pe</strong>line<br />

FIGURE 12-1 General schematic for tailings disposal pi<strong>pe</strong>line.<br />

spigot<br />

cyclones<br />

submerged<br />

disposal<br />

fines for<br />

submerged<br />

disposal<br />

coarse<br />

for banks<br />

of pond<br />

fines for<br />

submerged<br />

disposal<br />

coarse<br />

for banks<br />

of pond<br />

Tailings pond (single or multiple cells)<br />

(or sometimes the sea is used instead of man made ponds)<br />

Tailings Disposal Ponds (Dams)


Mineral<br />

Process<br />

Plant<br />

<strong>slurry</strong><br />

Field pump test<br />

Loop to adjust<br />

concentration<br />

thickener<br />

Emergency<br />

pond<br />

isolation<br />

valve<br />

clarified<br />

water<br />

dilution<br />

water<br />

corrosion<br />

inhibitors<br />

concentrate<br />

pi<strong>pe</strong>line<br />

feed & storage<br />

agitator tank<br />

pi<strong>pe</strong>line feed pump station<br />

centrifugal pumps up to 7.7 MPa,1100 psi<br />

reciprocating up to 18 MPa (2600psi))<br />

FIGURE 12-2 General schematic for concentrate pi<strong>pe</strong>line.<br />

concentrate<br />

pi<strong>pe</strong>line<br />

filter/dewatering<br />

plant


FEASIBILITY STUDY FOR A SLURRY PIPELINE<br />

12-2 RHEOLOGY, THICKENER<br />

PERFORMANCE, AND PIPELINE SIZING<br />

12.5<br />

Thickeners are often located at the starting point of the pi<strong>pe</strong>line. Thickeners are installed<br />

for a certain production capacity and can be modified for higher output through the use of<br />

flocculants.<br />

Certain slurries, particularly those rich in fines, silt, and clay, can prove troublesome<br />

for the thickeners. At concentrations in excess of 50–55% by weight, the presence of such<br />

fines could dramatically increase the viscosity and yield stress. This in turn could force<br />

higher power to be needed for pi<strong>pe</strong>line feed pumps, or could force the o<strong>pe</strong>rator to dilute<br />

the <strong>slurry</strong>.<br />

Pilot plant tests are highly recommended. In fact, in large mines, a local pump test<br />

loop is sometimes built at the location of the thickeners. The underflow or the concentrate<br />

is pum<strong>pe</strong>d through the test loop in order to measure the viscosity and pressure drop. The<br />

information from the test loop is then used to adjust the o<strong>pe</strong>ration of the main pi<strong>pe</strong>line<br />

pumps and to feed information to the dewatering plant.<br />

During the feasibility study, samples of the ore are sent to a rheology lab. Samples<br />

should be taken from different boreholes. Some boreholes may yield coarser material at<br />

the higher levels but finer materials at dee<strong>pe</strong>r depth. This information is used to predict<br />

the <strong>pe</strong>rformance of the pumps throughout the lifetime of the project. For example, in the<br />

earlier years of the project, the <strong>slurry</strong> may be coarser and of heterogeneous flow. As the<br />

life of the project progresses, finer material may be pum<strong>pe</strong>d at higher concentrations as<br />

non-Newtonian flows.<br />

Samples from different boreholes are also mixed for testing. The blended samples are<br />

quite important as the thickeners may be handling soils from different excavation points,<br />

such as a mixture of sulfides and oxides in different proportions.<br />

From the rheology of the <strong>slurry</strong> and the optimum <strong>pe</strong>rformance of the thickeners, the<br />

<strong>slurry</strong> engineer decides the range of concentrations needed to pump the <strong>slurry</strong>. The s<strong>pe</strong>ed<br />

of o<strong>pe</strong>ration is then decided on the basis of the ratio of coarse to fine particles, the velocity<br />

of deposition, and the friction losses.<br />

Example 12-1<br />

Samples of tailings from a cop<strong>pe</strong>r process plant are tested for viscosity and yield stress.<br />

Results are plotted in Table 12-1. Determine the maximum concentration for designing<br />

the thickness or pumping of <strong>slurry</strong><br />

It is obvious from the data that the viscosity and yield stress rise sharply above a<br />

weight concentration of 55%. The <strong>slurry</strong> engineer would be wise to consider o<strong>pe</strong>rations<br />

above 55% as unstable.<br />

Having decided that the maximum weight concentration is 55%, the information is<br />

then given to the process engineer in charge of selecting the thickener. Upon review of the<br />

TABLE 12-1 Combined Fine Tailings<br />

Weight concentration Reduced viscosity (<strong>slurry</strong>/water) Yield stress (Pa)<br />

40 4.95 1.5<br />

44 7.45 2.7<br />

49.9 12.5 6.2<br />

55 26.4 12.4<br />

60 45 25


12.6 CHAPTER TWELVE<br />

data, the process engineer notices that the thickeners may <strong>pe</strong>rform well without flocculants<br />

up to a maximum weight concentration of 50%. Because flocculants are ex<strong>pe</strong>nsive,<br />

a trade-off study is conducted on the power consumption of pumping <strong>slurry</strong> at 55%. The<br />

study reveals that there is an important increase in capital cost if investment in s<strong>pe</strong>cial<br />

thickeners is made to thicken the <strong>slurry</strong> at 55%. The amount of ex<strong>pe</strong>nsive flocculants for<br />

a weight concentration of 55% increases the o<strong>pe</strong>rating cost and the <strong>slurry</strong> is more viscous<br />

at 55% weight concentration. Despite the fact that there is an increase in the amount of<br />

water pum<strong>pe</strong>d at 50% weight concentration, a good compromise is found between cost of<br />

o<strong>pe</strong>ration of the thickeners and the capital costs needed for the pi<strong>pe</strong>line to handle the flow<br />

for o<strong>pe</strong>ration at a weight concentration of 53%. The thickeners, pi<strong>pe</strong>line size, and pumps<br />

are then sized to produce <strong>slurry</strong> at a 53% concentration by weight.<br />

Thickeners (Figure 12-3) are considered the starting point of tailings and concentrate<br />

<strong>slurry</strong> pi<strong>pe</strong>lines. For tailings pi<strong>pe</strong>lines, they feed directly into the tailings sump, but for<br />

concentrate pi<strong>pe</strong>lines they feed s<strong>pe</strong>cial storage tanks with agitators.<br />

The sump for the tailings pi<strong>pe</strong>line (Figure 12-4) may be built of concrete or rubberlined<br />

steel. A number of pi<strong>pe</strong>s are installed in the feed side such as:<br />

� Dilution process water pi<strong>pe</strong>s<br />

� <strong>slurry</strong> pi<strong>pe</strong>s<br />

� pi<strong>pe</strong>s from emergency ponds<br />

The concentrate storage tanks for <strong>slurry</strong> pi<strong>pe</strong>lines are essentially large tanks with agitators<br />

(Figure 12-5). A small pump test loop near these tanks is used to test the concentrate before<br />

feeding it to the pi<strong>pe</strong>line pumps. Feed is essentially from the thickeners, but continuous<br />

agitation in the tank and addition of viscosity control agents, corrosion inhibitors, and<br />

even some dilution water are part of the process.<br />

Not every <strong>slurry</strong> pi<strong>pe</strong>line requires thickeners. Dredging pi<strong>pe</strong>lines and phosphate rock<br />

pumping are both transport low-concentration slurries. These slurries are pum<strong>pe</strong>d over<br />

shorter distances and use pi<strong>pe</strong>s and pumps that are physically relocated from one point to<br />

another. Some pi<strong>pe</strong>lines o<strong>pe</strong>rate totally as an o<strong>pe</strong>n channel flow, such as the tailings<br />

pi<strong>pe</strong>line of Southern Peru Cop<strong>pe</strong>r in Peru.<br />

FIGURE 12-3 Thickeners. (Courtesy of Geho Pumps.)


FEASIBILITY STUDY FOR A SLURRY PIPELINE<br />

FIGURE 12-4 Sump for tailings pi<strong>pe</strong>line.<br />

FIGURE 12-5 Concentrate storage tank.<br />

12.7


12.8 CHAPTER TWELVE<br />

12-3 RECLAIM WATER PIPELINE<br />

Although <strong>slurry</strong> is pum<strong>pe</strong>d from the process plant to a tailings disposal site, reclaim water<br />

is often pum<strong>pe</strong>d back from the tailings pond to the mine. A popular method of feeding the<br />

reclaim water into a pi<strong>pe</strong>line is by installing vertical turbine (mixed flow) pumps on a<br />

barge or onshore near a pump station (Figure 12-6). The number of stages of these vertical<br />

pumps is set by the total dynamic head and the possibility of installing booster pump<br />

stations along the pi<strong>pe</strong>line route. The pi<strong>pe</strong>line material may be constructed of steel or<br />

high-density polyethylene. The latter, however, is limited to a pressure rating of 1.4 MPa<br />

(200 psi) on large pi<strong>pe</strong> sizes (see Chapter 2 for more details on the pressure rating of<br />

HDPE).<br />

If the reclaim water pi<strong>pe</strong>line is steel and the tailings have been neutralized for corrosion<br />

using lime, the pi<strong>pe</strong>line may gradually suffer from deposition of lime on the inside<br />

walls. Over time, this increases the pi<strong>pe</strong>’s roughness; friction losses increase and the<br />

<strong>pe</strong>nalty could be higher power consumption. To prevent such a problem, polypig launching<br />

and receiving stations are installed at the start and end of the pi<strong>pe</strong>line. Polypigs are<br />

sponge-filled bullets with brushes sized to the pi<strong>pe</strong> diameter. As they move in the pi<strong>pe</strong>,<br />

they clean its surface.<br />

The methods of sizing the reclaim pi<strong>pe</strong>line for single-phase water were covered in<br />

Chapter 2.<br />

Floating pump stations are often designed as a catamaran for adequate stability. The<br />

pumps are located in the middle of the barge. The catamaran is built with buoyancy<br />

tanks on each side that can be filled. Some catamarans have a false bottom to protect the<br />

suction of the pump. Reclaim water enters from the side pump inlet via a pro<strong>pe</strong>r fish<br />

screen. The fish screen prevents any fish or aquatic plants from being pum<strong>pe</strong>d back to<br />

the mine.<br />

FIGURE 12-6 Pump station.


12-4 EMERGENCY POND<br />

FEASIBILITY STUDY FOR A SLURRY PIPELINE<br />

12.9<br />

An emergency pond should be carved out or built at the start of the pi<strong>pe</strong>line near the<br />

pump station or at the lowest point in the pi<strong>pe</strong>line. The purpose of the emergency pond is<br />

to provide a means of draining the <strong>slurry</strong> pi<strong>pe</strong>line (Figure 12-7). The decision to dig an<br />

emergency pond is often based on the ability to restart a pi<strong>pe</strong>line after a shut down.<br />

Restarting is often difficult with particularly coarse slurries, taconite, sand, and dredging<br />

rocks. With finer slurries and clays, it may be possible to restart the pi<strong>pe</strong>line without<br />

draining it, provided that the maximum slo<strong>pe</strong> does not exceed the critical value (discussed<br />

in Chapter 4).<br />

An emergency pond is needed in cold areas to avoid freezing the pi<strong>pe</strong>line after a shutdown.<br />

Sometimes a s<strong>pe</strong>cial valve chamber is installed with a valve on a tee branch. This<br />

valve automatically o<strong>pe</strong>ns on power failure to divert <strong>slurry</strong> to the emergency pond.<br />

An emergency pond needs its own pumping system. It can consist of a vertical <strong>slurry</strong><br />

pump floating on a pontoon or barge (Figure 12-8). The sump pump feeds a booster pump<br />

that redirects the <strong>slurry</strong> back to the pi<strong>pe</strong>line pump box or back to the thickeners.<br />

Submersible pumps (Figure 12-9) with augers are also used for emergency ponds near<br />

thickeners, particularly with concentrate pi<strong>pe</strong>lines. S<strong>pe</strong>cial water sparges are installed<br />

FIGURE 12-7 Emergency pond.


floats<br />

pump<br />

FIGURE 12-8 Emergency pond pumping system.<br />

FIGURE 12-9 Submersible pump.<br />

12.10<br />

motor


around the emergency pond to dilute the <strong>slurry</strong>. Although fairly reliable, submersible<br />

pumps require a s<strong>pe</strong>cial shop to rebuild them and replace the seals. In remote mines, they<br />

tend to be less popular than vertical cantilever pumps.<br />

It is also recommended to install emergency ponds near booster pump stations. These<br />

should be connected to the booster station by a drainpi<strong>pe</strong>. A pontoon on the pond with a<br />

cantilever pump (Figure 12-8) is recommended to pump back the spill to the booster station<br />

pump box.<br />

If the tailings dam is to be located in a flood plain, the civil engineer may recommend<br />

an emergency spillway. A decant pond may serve as an emergency spillway. Sometimes<br />

it is more economical to provide adequate height of the walls of the dam to contain the<br />

1:100 year flood, particularly when purchasing more land is an ex<strong>pe</strong>nsive proposition.<br />

12-5 TAILINGS DAMS<br />

FEASIBILITY STUDY FOR A SLURRY PIPELINE<br />

Many pi<strong>pe</strong>lines are used for pumping tailings. Selecting a site for disposal of tailings is<br />

based on many factors:<br />

� The tailings dam must be able to be used for the life term of the mine (e.g., 10–20<br />

years).<br />

� The site bedrock or foundation must be stable to build the dam walls. These are typically<br />

made of sand and coarse rejects and some are built at a rate of 4.6 m (15 ft) <strong>pe</strong>r<br />

year.<br />

� The site must not interfere with future expansion of the mine and must not be on an ore<br />

deposit. For this reason, the tailings disposal system is sometimes a long distance away<br />

from the mine or surrounding economic centers (towns, cities, and agricultural fields).<br />

� The tailings disposal area must be designed to minimize contamination such as seepage<br />

of liquids to surrounding areas.<br />

� The volume of the tailings containment must be calculated to account for disposal volumes,<br />

runoff of snow or rain, and the pumping out of reclaim water.<br />

� The process of separation of slimes from coarse materials at the tailings dam must be<br />

designed carefully. It can be as simple as a spigot when there is a considerable portion<br />

of coarse materials, or as complex as a two-stage cyclone when there are a lot of fines<br />

in the tailings.<br />

� Accessibility to the site is important for repairs and for construction of the tailings<br />

dam.<br />

The guidelines for constructing a tailings dam have been established by the International<br />

Commission on Large Dams (1982). These are reviewed briefly in the following<br />

paragraphs. These are general principles that must be adapted to every site and condition.<br />

It is important to be able to separate the coarse from the fine particles when building a<br />

tailings dam. The coarse solids are used to build the dam walls, whereas the fines are used<br />

to form the beaches (Figure 12-10).<br />

12-5-1 Wall Building by Spigotting<br />

12.11<br />

One method of constructing the walls of a dam is to use the coarse material in the tailings.<br />

The fines or slimes are allowed to sink to the bottom of the tailings pond or to form


12.12 CHAPTER TWELVE<br />

toe<br />

trench<br />

Rock<br />

Toe<br />

<strong>slurry</strong><br />

(coarse material)<br />

beaches between the water and the dam. When there is a high content of coarse particles,<br />

as in the taconite mines of Minnesota (U.S.A.), it is sufficient to use a spigot to separate<br />

coarse from fine particles. At the exit from the spigot, the coarse particles separate under<br />

gravity and pressure while the finer particles are carried further away. A bulldozer relocates<br />

and then compacts the material by rolling over it. The banks are gradually built this<br />

way. This o<strong>pe</strong>ration is difficult in the winter in Canada, Siberia, and northern United<br />

States. Therefore, the actual construction of the dam is limited to the summer months.<br />

Although the great majority of tailing dams are built on the concept of a single spigot,<br />

some use the concept of multiple parallel spigots. In the single-spigot approach, all the<br />

tailings are dis<strong>pe</strong>nsed at one point. After a couple of days or so, the spigot is then moved<br />

approximately 15 m (50 ft) away. At each location, the bulldozer is brought in to compact<br />

the coarse material. The banks of the dam are thus gradually built.<br />

In the multispigot system, the spigots are fixed in place. The diameter of the pi<strong>pe</strong>line<br />

is gradually reduced around the tailings dam. This method is particularly interesting in<br />

very cold climates when construction of the dam is difficult in the six months of the year<br />

when construction is not possible.<br />

12-5-2 Deposition by Cycloning<br />

cyclone<br />

underflow<br />

filter drain bed<br />

cyclone overflow<br />

is used to make<br />

beaches of fines<br />

beaches of fines<br />

FIGURE 12-10 Using tailings to build dams and beaches.<br />

decant water intake<br />

A spigot may not be sufficient to separate coarse from fine particles. More pressure and<br />

force may be needed. One particularly useful piece of equipment is the hydrocyclone<br />

(which was presented in Chapter 7). The coarse material is diverted to the underflow and<br />

the finer material to the overflow. In a certain ratio of coarse and fine particles, a single<br />

cyclone is sufficient, but when the coarse material is less than 25% of the tailings, two cyclones<br />

in series may be needed. It is strongly recommend that a cyclonability test be conducted<br />

in a lab before deciding whether a single-stage or a two-stage cyclone is needed.<br />

Example 12-2<br />

Tailings from a mine were tested for cyclonability. The following particle size distribution<br />

was obtained:<br />

Particle size (microns) 152 110 74 53 44 37 29 25 22 17<br />

Cumulative % passing 83.3 75.4 67 61 54 52 50 44 42 40<br />

pond


It is clear that the fines (


12.14 CHAPTER TWELVE<br />

Rock Toe<br />

Toe<br />

Trench<br />

Coarse material<br />

used for lifts<br />

1<br />

12-5-2-1 Mobile Cycloning by the Upstream Method<br />

When there is not a sufficient <strong>pe</strong>rcentage of coarse material in the tailings, it is very difficult<br />

to build a high dam. The approach is then to move the crest of the dam progressively<br />

upstream and to fill more and more surface area. This method is called the “upstream<br />

technique” (Figure 12-11). In this technique, the delivery pi<strong>pe</strong> is initially installed at the<br />

toe wall of the dam. The cyclone underflow is placed inside the toe wall and therefore upstream<br />

of the pi<strong>pe</strong> and partially on top of the fines beach.<br />

12-5-2-2 Mobile Cycloning by the Downstream Method<br />

The <strong>slurry</strong> pi<strong>pe</strong>line is placed on the internal dividing wall between the fines beach and the<br />

coarse material. The cyclone underflow material is placed between the toe and starter<br />

walls, or downstream from the delivery pi<strong>pe</strong>. The starter wall acts initially as a storage<br />

dam for fines until the coarse material develops an impoundment area to confine the fines.<br />

The crest of the dam moves progressively downstream as the impoundment area is<br />

raised. This provides a “full wedge” impoundment material of high strength. A bulldozer<br />

is used to push the coarse material upward and toward the toe of the dam. As the dam rises,<br />

the centerline of the impoundment area moves outward toward the outer toe of the<br />

dam (Figure 12-12).<br />

Rock<br />

Toe<br />

toe<br />

trench<br />

FIGURE 12-11 Upstream technique of dam building.<br />

Coarse<br />

material<br />

FIGURE 12-12 Downstream technique of dam building.<br />

2<br />

Crest of the dam<br />

2<br />

Annual<br />

lifts<br />

Dividing<br />

wall<br />

beaches of fines<br />

beaches of fines<br />

1


FEASIBILITY STUDY FOR A SLURRY PIPELINE<br />

FIGURE 12-13 Outside wall of tailings dam.<br />

Figure 12-13 shows the outside wall of a tailings dam. The steps represent successive<br />

lifts. Many tailing dams are constructed by annual lifts of 2.4 to 4.5 m (8 to 15 ft).<br />

Sometimes cycloning is done alternatively upstream and downstream to suit the split<br />

of coarse and fine materials. This technique is used when there is not sufficient coarse<br />

material for the downstream method, but there is sufficient material to create a wedge of<br />

coarse material that could be used more satisfactorily than a full upstream method (Figure<br />

12-14).<br />

12-5-2-3 Deposition by Centerline<br />

In this method, the crest of the dam is fixed in plan with res<strong>pe</strong>ct to the toe wall as the level<br />

of the impoundment is raised.<br />

12-5-2-4 Multicellular Construction<br />

One method of constructing a dam consists of dividing the impoundment area into a number<br />

of cells or small lakes and filling them one at a time. If one of them is used as a source<br />

of reclaim water, it may be set at a lower elevation and therefore fed by the decant water.<br />

This method is popular in hilly regions were it is difficult to find flat ground and the cost<br />

of excavating a large pond would be prohibitive.<br />

12-6 SUBMERGED DISPOSAL<br />

12.15<br />

When the material is simply too fine to build a dam, it may be recommended to submerge<br />

the disposal point. When the volume of the tailings is also too low, as in certain gold


12.16<br />

coarse<br />

by<br />

downstream<br />

method<br />

toe<br />

drain<br />

Rock<br />

toe<br />

FIGURE 12-14 Alternate upstream and downstream technique of dam building.<br />

4<br />

2<br />

5<br />

3<br />

1<br />

Starter<br />

wall<br />

coarse<br />

by<br />

upstream<br />

method<br />

beaches<br />

of<br />

slimes


FEASIBILITY STUDY FOR A SLURRY PIPELINE<br />

mines where tailings may be pum<strong>pe</strong>d at a rate of 20 m 3 /hr (88 US gpm), submerged disposal<br />

in a lake is an acceptable method.<br />

12-6-1 Sub-Sea Deposition Techniques<br />

The deposition of city wastewater in the sea has been done by large cities for a number of<br />

decades. Since the late 1960s, more and more mines have turned to the sea for disposal of<br />

tailings. In Peru, Southern Peru Cop<strong>pe</strong>r has been disposing its tailings in the sea since the<br />

early 1970s. In Chile, CMP has been disposing hematite tailings in the sea since the early<br />

1990s. The disposal of tailings in the sea must take into account low and high tides, the<br />

sea currents, and the redistribution of tailings. It must be environmentally friendly. Over<br />

the years, new beaches may form from accumulated solids and new disposal points must<br />

be selected.<br />

12-7 TAILINGS DAM DESIGN<br />

12.17<br />

Having determined the location of the tailings dam, and having determined its geometry<br />

and the rheology of the tailings, there are still a number of factors that affect the dam design.<br />

Topsoil and unsuitable foundation materials may need to be removed within the footprints<br />

of the dam. This material may have to be stockpiled for eventual closure and reclamation<br />

at the end of the project. A high starter dam may have to be built initially from<br />

compacted waste rock hauled from the mine.<br />

It may be necessary to install a blanket drain of high-<strong>pe</strong>rmeability material. A geotextile<br />

filter blanket may have to be installed to separate the waste rock from cycloned sand.<br />

To monitor the structural integrity of the dam, vibrating wire piezometers (VWPs) may be<br />

installed beneath the foundation of the dam. The VWP should be placed in the blanket<br />

drain, within the starter dam and within the cycloned sands.<br />

A checklist for the design of a site tailings dam is presented in Table 12-2.<br />

TABLE 12-2 Checklist for Site Tailings Dam Design Information<br />

Crest elevation of starter dam<br />

Final crest elevation of the dam<br />

Crest width<br />

Maximum crest length<br />

Average upstream slo<strong>pe</strong> of starter dam<br />

Average downstream slo<strong>pe</strong> of starter dam<br />

Average upstream slo<strong>pe</strong> of cycloned sand<br />

Average downstream slo<strong>pe</strong> of cycloned sand<br />

Blanket drain


12.18 CHAPTER TWELVE<br />

12-8 SEEPAGE ANALYSIS OF<br />

TAILINGS DAMS<br />

In order to determine the ex<strong>pe</strong>cted water level in the completed dam, seepage analysis is<br />

conducted using the data from soils encountered at the site, as well as soils to be part of<br />

the tailings or construction. Seepage from the blanket drain or cycloned sands should be<br />

collected in a trench and sent to a sump where it can be measured.<br />

Seepage analysis is not the responsibility of the <strong>slurry</strong> engineer and requires a good<br />

geologist and complex computer programs. Seepage analysis is usually based on the <strong>pe</strong>rmeability<br />

of the soils as shown in Table 12-3. It is discussed here so that the <strong>slurry</strong> engineer<br />

may design an appropriate monitoring sump.<br />

12-9 STABILITY ANALYSIS FOR<br />

TAILINGS DAMS<br />

A stability analysis must be conducted to find the critical failure surface. This helps determine<br />

the stable slo<strong>pe</strong>. Stability charts are available from the U.S. Naval FEC Soils Mechanics<br />

Design Manual 7.01 (1986). Different commercial software packages are available.<br />

They use data from bore hole and test pit logs.<br />

The data from the stability analysis is critical to determine the final height or surface<br />

area of the dam. Obviously, if the soils are weak and unstable, it will not be possible to<br />

build a high dam. More surface area will be required with longer pi<strong>pe</strong>s.<br />

The stability analysis includes determining the angle of friction for sands, the cohesion<br />

strength for clays and silts, as well as the density of soils. These determine the strength of<br />

the soils. Static and dynamic slo<strong>pe</strong> stability analyses, along with seismic analysis must<br />

also be <strong>pe</strong>rformed.<br />

Settlement may occur during the construction phase of the dam or during its o<strong>pe</strong>ration.<br />

Alluvial soils under tailings dams are prone to settlement up to a depth of 3 m (9.85 ft).<br />

An ex<strong>pe</strong>rt should be involved in the design of the walls.<br />

The risk of liquefaction needs to be assessed, particularly when the tailings have a<br />

very large <strong>pe</strong>rcentage of fines. The plasticity index (see Chapter 1) needs to be measured<br />

to assess the role of liquefaction.<br />

When tailings dams are built on stable foundations, they can be built to a height of 60<br />

m (200 ft). But if an accident hap<strong>pe</strong>ns, they may spill harmful liquids such as cyanide solutions,<br />

with disastrous consequences.<br />

TABLE 12-3 Examples of Materials Permeability for Seepage Analysis<br />

Material Permeability (m/s)<br />

Foundation clay soils 8.4 × 10 –9<br />

Foundation silt 2.5 × 10 –7<br />

Foundation gravel soil 10 –4<br />

Foundation waste rock and sand for starter dams 5 × 10 –5<br />

Foundation soils 4 × 10 –6<br />

Impoundment tailings 6 × 10 –8<br />

Drain sand 10 –4<br />

Cycloned embankment sand 4 × 10 –6


12-10 EROSION AND CORROSION<br />

Erosion and corrosion data are very important at an early stage of the design. There are a<br />

number of recommended tests. One method is the ASTM G75 Slurry Abrasivity Determination<br />

by the Miller number system. The Miller number is established as the relative rate<br />

of mass or volume loss in 2 hours. Slurries with Miller numbers smaller than 50 are considered<br />

relatively nonabrasive and can be pum<strong>pe</strong>d with piston reciprocating pumps. Slurries<br />

with Miller numbers above 80 are considered relatively abrasive and require flushed<br />

plunger or membrane pumps (see Section 10-6).<br />

One method to reduce corrosion of steel pi<strong>pe</strong>s is to raise the pH by adding lime. The<br />

engineer also has a number of choices for pi<strong>pe</strong>line materials:<br />

� For applications up to 1400 kPa (200 psi), high-density polyethylene pi<strong>pe</strong>s are available<br />

for a number of low to medium abrasive slurries.<br />

� For more abrasive slurries with particles up to 6 mm ( 1 – 4 in), rubber-lined pi<strong>pe</strong>s are<br />

available for pressures up to 1200 psi.<br />

� For coarser slurries and dredging applications, plain steel pi<strong>pe</strong>s are available.<br />

Unlined plain steel pi<strong>pe</strong>s are sometimes installed with sacrificial thickness. This is<br />

particularly the case with continuous welded underground pi<strong>pe</strong>lines. The reader should<br />

refer to Chapters 9 and 11 for some of the particular approaches used to select materials<br />

for different slurries. Some slurries such as laterite and taconite has been found to be very<br />

abrasive with HDPE pi<strong>pe</strong>s.<br />

The actual design and construction of the pi<strong>pe</strong>line should comply with ANSI/ASME<br />

B31.11 (1989 Edition) Slurry Transportation Piping Systems (discussed in the previous<br />

chapter), as well as the standards of the American Water Works Association and local and<br />

national regulations.<br />

12-11 HYDRAULICS<br />

FEASIBILITY STUDY FOR A SLURRY PIPELINE<br />

To conduct a pro<strong>pe</strong>r study on friction losses, it is a good idea to begin with practical historical<br />

cases, as discussed in Chapter 11, as well as test data. The methods for hydraulic<br />

friction loss estimation have been covered extensively in Chapters 2 to 5 for closed conduits<br />

and in Chapter 6 for o<strong>pe</strong>n channel flow. A pi<strong>pe</strong>line may consist of both ty<strong>pe</strong>s of<br />

flow. For example, <strong>slurry</strong> may be pum<strong>pe</strong>d to a tank situated at a high point, and then be<br />

allowed to flow by gravity to the tailings pond.<br />

12-12 PUMP STATION DESIGN<br />

12.19<br />

In Chapter 8, it was indicated that the maximum head from rubber-lined pumps is about<br />

30 m (100 ft) and about 55 m (180 ft) for all-metal pumps. This rule of thumb quickly<br />

helps the engineer determine the minimum number of pumps to be installed in series.<br />

Example 12-3<br />

A <strong>slurry</strong> pi<strong>pe</strong>line is designed to handle flow at a pressure of 3.5 MPa (� 500 psi). The<br />

<strong>slurry</strong> s<strong>pe</strong>cific gravity is 1.5. Determine the number of pumps to be installed in the pump<br />

station (assume a head factor of 0.83).


12.20 CHAPTER TWELVE<br />

Solution<br />

The total dynamic head is first calculated:<br />

TDH = P/(�g) = 3.5 × 106 /(1500 × 9.81) � 231 m (or 758 ft)<br />

Assuming a head ratio factor of 0.83 due to the concentration of <strong>slurry</strong>, the corrected head<br />

is 231/0.83 = 279 m (or 915 ft).<br />

Iteration 1<br />

Assuming all-metal construction at 55 m/stage (180 ft), this would be closer to 5.1 stages.<br />

The engineers must therefore install 6 stages:<br />

279/6 = 46.5 m/stage<br />

For the sake of process control, the engineer may install the variable frequency drive<br />

in the last stage to accumulate fluctuation of (15% of head or ±7m (21 ft)—46.5 – 7 =<br />

39.5 m (129.6 ft) on the lower side for the last stage and 46.5 + 7 = 53.5 m (176 ft) on the<br />

up<strong>pe</strong>r side of the last stage. This leaves the opportunity to use fixed-s<strong>pe</strong>ed motors on the<br />

first 5 stages, and a variable frequency drive for the final stage.<br />

If rubber-lined pumps are to be used, and assuming about 30 m/stage, 10 pumps are<br />

required. To cut down on the number of rubber-lined pumps, polyurethane-lined pumps<br />

may be used because they can o<strong>pe</strong>rate up to a tip s<strong>pe</strong>ed of 32 m/s (6300 ft/min). Considering<br />

that the head is proportional to the square of the tip s<strong>pe</strong>ed:<br />

(32/28) 2 × 30 � 39 m/stage (128 ft/stage)<br />

279 m/39 m � 7.2 m/stage or 8 pumps in a series<br />

If rubber-lines pumps are to be used, and assuming about 30 m/stage, this results in an<br />

installation requiring 10 pumps.<br />

12-13 ELECTRIC POWER SYSTEM<br />

It would be beyond the sco<strong>pe</strong> of this book to discuss electrical engineering. However,<br />

some basic concepts need to be reviewed. The electric starters and transformers should be<br />

installed above the 1:100 flood level or above the estimated <strong>slurry</strong> level in case of flooding<br />

due to a pi<strong>pe</strong>line rupture or back flow.<br />

The estimated consumed power of reclaim water pumps and booster <strong>slurry</strong> pumps is<br />

supplied to the electrical engineer. Power is brought to the mining site using 10 kV or<br />

14.6 kV overhead or underground lines. Overhead transmission lines may be of an aluminum<br />

conductor, steel reinforced (ACSR). Underground cables may be made of aluminum<br />

or cop<strong>pe</strong>r. On-site power generation is sometimes considered.<br />

Overhead lines are less reliable than underground lines as they are subject to the following<br />

risks:<br />

� Trees add to the risks of overhead lines because the branches can sway or fall onto the<br />

line, causing the line to fault.<br />

� Lightening storms in some areas require the lines to have s<strong>pe</strong>cial lightening protection.<br />

� Wind and ice formation can also cause momentary line-to-line shorts or even failure.<br />

Although underground lines enjoy the protection of the earth overfill, they are not<br />

problem-free. They are more ex<strong>pe</strong>nsive to install and suffer from longer mean time to repair<br />

(MTTR) than overhead lines.


FEASIBILITY STUDY FOR A SLURRY PIPELINE<br />

Poles for overhead lines may be constructed of steel, concrete, or wood. Concrete is<br />

usually less ex<strong>pe</strong>nsive than steel and more resistant than wood.<br />

The line size is optimized to minimize o<strong>pe</strong>rating resistance and cost over the life of the<br />

project.<br />

12-14 TELECOMMUNICATIONS<br />

Along the corridor of the power line, a system control and data acquisition (SCADA) or<br />

instrumentation line is run back to the main control station. Telephone wires or fiber optics<br />

are used to monitor remote start or shut pumps and valves. The computer system that<br />

monitors the local control may be “programmable local controller” (PLC) based. The centralized<br />

control room is usually called the distributed control system (DCS). SCADA controlled<br />

<strong>systems</strong> are designed to include:<br />

� Automatic controls to o<strong>pe</strong>rate as required by the process<br />

� Manual controls to override SCADA locally<br />

� Remote control for remote start/stop or o<strong>pe</strong>n/close as required<br />

Motor protection devices should interface to the SCADA at least for status display.<br />

Motor control circuits must be hardwired for motor start/stop control and run status.<br />

“Four wire” ty<strong>pe</strong>s of motor control require one normally o<strong>pe</strong>n contact to start the motor<br />

and one normally closed to stop it. The advantage is that the motor will maintain current<br />

o<strong>pe</strong>rating state (running or stop<strong>pe</strong>d) in case of PLC failure or accidental shut-down.<br />

Indoor enclosures for control <strong>systems</strong> in North America should be NEMA 12 with<br />

locking handles. Outdoor enclosures must be NEMA 4X with locking handles.<br />

It is recommended that each PLC have at least three communication ports—one for<br />

the local o<strong>pe</strong>rator interface, another for the host communications, and a third for the programming<br />

terminal. It is recommended to use modular input/output cards. The system<br />

should be designed with 20% spare capacity for input/output and with the capability for<br />

100% expansion.<br />

12-15 TAILINGS DAM MONITORING<br />

12.21<br />

Monitoring tailings dams can prevent dangerous accidents. A number of methods are recommended:<br />

� Daily visual ins<strong>pe</strong>ctions. Records should be kept.<br />

� Measurement of dry density and moisture contents to ASTM D 3017 standard every<br />

800 cubic meters (28,250 cubic feet) of placed cycloned sand.<br />

� Measurement of dry density and moisture to ASTM D 1556 every three months.<br />

� Determining grain size distribution to ASTM D 422 every 5000 cubic meters (or<br />

176,570 cubic feet) of placed cycloned sand, or after each shut-down of the cyclone<br />

station, or after important changes to the ore.<br />

� Measurement of s<strong>pe</strong>cific gravity of the tailings in accordance with ASTM D 854, on a<br />

monthly basis, after a shut-down of the cyclone station, or significant change to the<br />

ore.


12.22 CHAPTER TWELVE<br />

� Monthly measurements of the compaction of the deposited sands in accordance with<br />

ASTM D 698 for relative density.<br />

� Measurement of the shear strength characteristics to ASTM D 4767 every 100,000 cubic<br />

meters (3.5 million cubic feet) of placed cycloned sand.<br />

� Weekly measurement of pore-water pressure by the vibrating wire piezometer method.<br />

� Monthly surface survey taken every 10 m (30 ft) along the downstream slo<strong>pe</strong> and starting<br />

from the crest.<br />

� Weekly monitoring of seepage by using a notch weir.<br />

The results need to be compared with the original design calculations and assumptions to<br />

make appropriate decisions.<br />

12-16 CHOKE STATIONS AND IMPACTORS<br />

Chokes and choke stations are used to maintain a full pi<strong>pe</strong>line by applying an artificial<br />

pressure loss at the end of the line or at a particular point. Chokes may be pinch valves or<br />

cermet nozzles. Cermet nozzles are made from a composite of metal and ceramic.<br />

Impactors (Figure 12-15) are essentially spring actuated relief valves that protect the<br />

pi<strong>pe</strong>line from surges and water-hammer-related damage. Impactors were develo<strong>pe</strong>d for<br />

use in the phosphate mines of Florida (U.S.A.), and are gaining acceptance in other fields<br />

of mining. S<strong>pe</strong>cial air relief valves (Figure 12-16) may need to be installed on tailings and<br />

<strong>slurry</strong> pi<strong>pe</strong>lines at high points to protect the <strong>slurry</strong> <strong>systems</strong>. For <strong>slurry</strong>, the long body design,<br />

familiar in sewage pi<strong>pe</strong>lines, is recommended.<br />

12-17 ESTABLISHING AN APPROACH FOR<br />

START-UP AND SHUTDOWN<br />

It is very important to avoid blocking <strong>slurry</strong> lines. The profile of the pi<strong>pe</strong>line must not<br />

include any steep gradients in excess of the critical slo<strong>pe</strong>. Slurry pi<strong>pe</strong>lines should be<br />

FIGURE 12-15 Impactor.


FEASIBILITY STUDY FOR A SLURRY PIPELINE<br />

FIGURE 12-16 Air relief valve.<br />

purged after they are shut down in order to avoid sanding the lines. It is not recommended<br />

to start empty lines with <strong>slurry</strong>. It is preferable to first pump water or thickener<br />

overflow into the line before gradually introducing <strong>slurry</strong>. Some <strong>slurry</strong> lines also require<br />

minimum static head or pressure at start-up. Filling the line with water creates the<br />

pressure needed for start-up, and is more effective than pumping against the air in an<br />

empty line.<br />

12-18 CLOSURE AND RECLAMATION PLAN<br />

12.23<br />

It is recommended even at the very start of a feasibility study or basic engineering exercise<br />

to have a closure and reclamation plan. This is really the work of the environmental<br />

engineer, civil engineer, and geophysicist. The hidden costs of abandoning a tailings dam<br />

can be horrendous.<br />

The closure plan may involve planting trees, adding topsoil on the impoundment, or<br />

backfilling and grading with the use of tracked dozers. In one interesting site in South<br />

America, a mine was disposing tailings in the sea. As the sand filled the shore, the seawater<br />

retracted and farmers started moving in and reclaiming the land for agriculture. The<br />

mine was far from reaching a point of closure, and had not established a reclamation plan.<br />

Although this was a convenient closure plan for the mine, it was not an environmentally


12.24 CHAPTER TWELVE<br />

safe solution. Tailings may contain chemicals that are particularly unhealthy to absorb, so<br />

it is safer to grow vegetation such as timber for industrial use than it is to grow plants for<br />

human or animal consumption. The <strong>slurry</strong> engineer must therefore work in close collaboration<br />

with an environmental engineer and oceanographer.<br />

12-19 ACCESS AND SERVICE ROADS<br />

Access and service roads are an important cost to factor into the budget. It is of utmost<br />

importance to be able to reach each pump station, valve chamber, choke station, and<br />

emergency pond and to be able to drive around a tailings dam. In large mines, these<br />

access roads are sometimes 20 m (60 ft) wide in order to accommodate heavy machinery.<br />

12-20 COST ESTIMATES<br />

Once the basic engineering scheme is completed, or even during its progression, data is<br />

gathered for cost estimates. Written quotes for pumps, equipment, valves, and cyclones<br />

are obtained from manufacturers. The contractors are requested to submit quotes for the<br />

pi<strong>pe</strong>line based on surveys and contour maps. Prices for earthworks are calculated on a<br />

cost-efficient method or using excavation, load, haul, placement, and compaction fleet. It<br />

is very useful to use local costs for similar projects because the labor costs, customs and<br />

duties, and equipment leases change from country to country.<br />

The s<strong>pe</strong>cific soil investigation along the footprint of the dam and along the pi<strong>pe</strong>line,<br />

may yield large quantities of topsoil that must be strip<strong>pe</strong>d away and stockpiled. This information<br />

should then be supplied to the contractor.<br />

Contractor overhead, profit, and mobilization and demobilization costs are calculated<br />

as an acceptable <strong>pe</strong>rcentage of the lump sum of all installation and construction costs. Often,<br />

the cost runs between 28–32% of the cost of the project. There may also be additional<br />

import duties applied to equipment and materials from outside the country or even inside<br />

the county. In some jurisdictions, these taxes are as high as 14–18%.<br />

12-20-1 Capital Costs<br />

Capital costs for a tailings project include the following (Table 12-4):<br />

� Capital submission costs, such as all project construction and management costs<br />

� Ongoing costs of studies, engineering, investigation, and new land purchases<br />

� Sunk costs, such as costs for project evaluation and investigation of already purchased<br />

land<br />

� Sustaining capital or future capital costs to extend blanket drains, starter dams, and access<br />

roads<br />

� Exit costs to close the project, revegetate, and move back stockpiled topsoil<br />

In some other applications, there may be a separate dewatering and filtering plant as<br />

well as a ship loading facility for a concentrate. Construction of such facilities may repre-


FEASIBILITY STUDY FOR A SLURRY PIPELINE<br />

TABLE 12-4 Checklist for Capital Costs Estimate for a Tailings Pi<strong>pe</strong>line<br />

Item Value in accepted currency<br />

1) Capital submission<br />

a) Access roads<br />

b) Contingency<br />

c) Thickeners modifications<br />

d) Cycloning or spigotting<br />

e) Impoundment<br />

f) Project management (including commissioning)<br />

g) Pi<strong>pe</strong>lines (<strong>slurry</strong> and water)<br />

h) Pump stations<br />

2) Ongoing costs<br />

a) Land purchase<br />

b) Project evaluation<br />

3) Sunk costs<br />

a) Previously purchased land<br />

b) Project evaluation<br />

Total capital submission<br />

Total ongoing costs<br />

Total ongoing costs<br />

Sustaining capital costs<br />

Exit capital costs<br />

sent an important capital investment involving the purchase of land and equipment, construction<br />

costs, power generation, mobilization, and demobilization.<br />

12-20-2 O<strong>pe</strong>ration Cost Estimates<br />

The annual cost of o<strong>pe</strong>rating a <strong>slurry</strong> pi<strong>pe</strong>line bears an important influence on the final<br />

designs. In the 1950s and 1960s, it was common to see many booster stations of two or<br />

three pumps in series. They were ex<strong>pe</strong>nsive to o<strong>pe</strong>rate and maintain. With the advent of<br />

high-pressure pumps up to 6.3 MPa (900 psi), the tendency is to centralize pump stations<br />

and eliminate some overhead lines, transformers, and trucks to move spare parts, etc.<br />

Annual o<strong>pe</strong>rating costs for a tailings system may include the following (Table 12-5):<br />

� Earthworks to grade and place sands<br />

� Costs of moving spigots, discharge pi<strong>pe</strong>lines, and cyclone underflow pi<strong>pe</strong>s<br />

12.25


12.26 CHAPTER TWELVE<br />

TABLE 12-5 Checklist for O<strong>pe</strong>ration Costs<br />

Activity Cost in approved currency<br />

Equipment, fuel, earthworks and piping movement<br />

Impoundment monitoring<br />

Labor<br />

Power requirement for <strong>slurry</strong> pumps<br />

Power requirement for reclaim pumps<br />

Cyclone data on power consumption<br />

Flocculants for thickeners<br />

Power transmission losses<br />

Maintenance and spare parts<br />

Maintenance contracts<br />

Others<br />

Other dewatering costs<br />

Ship loading costs for concentrations<br />

Contingency (10%)<br />

� Salaries for employees to manage and monitor the tailings impoundment area<br />

� Spare parts for pumps and cyclones and regular maintenance of mechanical and electrical<br />

equipment<br />

� Power for the tailings pumps<br />

� Power for the reclaim pumps<br />

� Cost of flocculants for the thickeners to maintain the concentration of the <strong>slurry</strong><br />

� Power transmission loss<br />

� Cost of relining the tailing pi<strong>pe</strong>s with rubber-lined cases<br />

� Cost of pi<strong>pe</strong>s and materials<br />

� Contract services to maintain pumps and equipment<br />

� Security and monitoring costs<br />

� Travel and training costs<br />

� Contingency costs (10%)<br />

There may be more cost studies needed to conduct to take in account variations in<br />

cost of material and inflation rate. These studies are better left to the ex<strong>pe</strong>rienced estimator.<br />

Despite the most detailed financial studies, natural disasters can wreak havoc with the<br />

most precise engineering estimates. For example, a landslide in Argentina buried a section<br />

of the concentrate pi<strong>pe</strong>line of Bajo Alumbera during the construction phase. It was<br />

necessary to reroute the pi<strong>pe</strong>line and add a booster pump station with a very ex<strong>pe</strong>nsive


price tag. The cost estimate is the result of teamwork between s<strong>pe</strong>cialists of various<br />

branches of engineering.<br />

12-21 PROJECT IMPLEMENTATION PLAN<br />

During the phase of basic engineering and feasibility studies, it becomes clear that there<br />

may be long delays in the purchase of equipment, mobilization of the construction fleet,<br />

obtaining <strong>pe</strong>rmits approvals, securing financing, etc. A list of tasks is established with estimates<br />

of required time to complete each step of construction and installation of the<br />

pi<strong>pe</strong>line and equipment. Such a list may include the following:<br />

� Development of a tailings management strategy to support the mining plan<br />

� Basic engineering designs and completion of prefeasibility studies<br />

� Approval of the prefeasibility studies by the owner and financiers<br />

� Detailed engineering designs and completion of feasibility study<br />

� Further approvals and <strong>pe</strong>rmits for construction<br />

� Land purchase and acquisition of rights of way or lease agreements on land<br />

� Detailed engineering and procurement<br />

� Construction of access roads, site preparation, blanket drain, foundation for thickeners,<br />

pump stations, and starter dams for tailings<br />

� Excavation for buried pi<strong>pe</strong>lines, or installation of pi<strong>pe</strong>lines and their supports<br />

� Installation of poles for overhead power lines or excavation for underground lines<br />

� Construction of pump stations, MCC buildings, and installation of equipment<br />

� SCADA (<strong>systems</strong> control and data acquisition) system installation<br />

� Start-up and commissioning<br />

� Demobilization of construction fleet<br />

� O<strong>pe</strong>ration and monitoring<br />

� Closure and reclamation<br />

Based on such a task list, the project milestones are established with a precise date set<br />

for start and completion of each phase of the project. To each milestone, a part of the<br />

budget is allocated.<br />

12-22 CONCLUSION<br />

FEASIBILITY STUDY FOR A SLURRY PIPELINE<br />

12.27<br />

The different phases of the engineering of a <strong>slurry</strong> pi<strong>pe</strong>line are presented in this chapter. It<br />

is not always a straightforward process and may involve trade-off studies based on the<br />

rheology of the <strong>slurry</strong>, the budget restrictions, the size and capabilities of the equipment,<br />

and mining plans.<br />

This effort is only accomplished through teamwork that involves engineers and designers<br />

from different professions including an estimator, a purchasing officer, technicians<br />

in test labs, geophysicists, and even s<strong>pe</strong>cialists for reclamation of plants and vegeta-


12.28 CHAPTER TWELVE<br />

tion. It is a very important step to sell the concept to decision makers and financing institutions,<br />

and the ex<strong>pe</strong>rt in <strong>slurry</strong> <strong>systems</strong> should be consulted on the various options.<br />

12-23 REFERENCES<br />

The International Commission on Large Dams. 1982. Manual on Tailings Dams and Dumps, Vol 45.<br />

Paris: Author.<br />

U.S. Naval FEC Soils Mechanics Design Manual 7.01 (1986).<br />

Further reading:<br />

Aplin, C. L. and G. O. Argall. 1972. Tailing disposal today. In Proceedings of the First International<br />

Tailing Symposium. Tucson, AZ. San Francisco: Miller Freeman.


APPENDIX B<br />

UNITS OF MEASUREMENT<br />

Acceleration<br />

To convert from To Multiply by<br />

foot/seconds2 (ft/sec2 ) meter/seconds2 (m/s2 ) 0.3048<br />

meter/seconds2 (m/s2 ) foot/seconds2 (ft/sec2 Angular Momentum<br />

) 3.28084<br />

To convert from To Multiply by<br />

kilogram-meter2 /second pound-foot2 /second 23.7304<br />

(kg-m2/s)<br />

Angular S<strong>pe</strong>ed<br />

(lb-ft/sec)<br />

To convert from To Multiply by<br />

radian/second revolution <strong>pe</strong>r minute 9.5493<br />

revolution <strong>pe</strong>r minute<br />

Area<br />

radian/second 0.10472<br />

To convert from To Multiply by<br />

square millimetre ( mm2 ) square inch (in2 ) 0.00155<br />

square inch (in2 ) square millimetre ( mm2 ) 645.161<br />

meter2 (m2 ) yard2 (yd2 ) 1.19599<br />

yard2 (yd2 ) meter2 (m2 ) 0.83612<br />

meter2 (m2 ) foot2 (ft2 ) 10.7639<br />

foot2 (ft2 ) meter2 (m2 ) 0.09290<br />

hectare (ha)<br />

Density<br />

acre 2.47105<br />

To convert from To Multiply by<br />

gram/centimeter3 (g/cm3 ) kilogram/meter3 (kg/m3 ) 1,000<br />

lbm/in3 kilogram/meter3 (kg/m3 ) 27,679<br />

lbm/ft3 kilogram/meter3 (kg/m3 ) 16.0185<br />

kilogram/meter3 (kg/m3 ) lbm/ft3 0.062428<br />

slug/foot3 kg/m3 515.379<br />

lbm/ft3 slug/foot3 0.03106<br />

B.1


B.2 APPENDIX B<br />

Energy<br />

To convert from To Multiply by<br />

British Thermal Units ( BTU)<br />

(ISO/TC 12)<br />

Joules ( J) 1,055.06<br />

Joule (J) foot-pound force (ft-lbf) 0.737562<br />

calorie (mean) (ca) Joules (J) 4.19<br />

megajoule (MJ) hour-horsepower (hph) 0.3725<br />

megajoule (MJ)<br />

Force<br />

kilowatt-hour (kwh) 0.27778<br />

To convert from To Multiply by<br />

dyne (dy) Newton (N) 1 × 10 –5<br />

pound-force (avoirdupois) (lbf) Newton (N) 4.4482<br />

Newton (N) lbf 0.224809<br />

kip<br />

Heat coefficient and transfer<br />

Newton 4,448.221<br />

To convert from To Multiply by<br />

Watt/meter2-degree Celsius British Thermal Unit/foot2- 0.17611<br />

(W/m2 oC) hour-degree Fahrenheit<br />

(BTU/ft2-hr o Length<br />

F)<br />

To convert from To Multiply by<br />

meter (m) foot (ft) 3.28084<br />

micrometer (�m) meter (m) 1 × 10 –6<br />

micrometer (�m) inch (in) 39.37 × 10 –6<br />

millimeter (mm) meter (m) 0.001<br />

millimeter (mm) inch (in) 0.03937<br />

fathom meter (m) 1.8288<br />

foot (ft) meter (m) 0.3048<br />

inch (in) meter (m) 0.0254<br />

inch (in) millimeter (mm) 25.4<br />

yard meter (m) 0.9144<br />

kilometer statute mile (US) 0.621371<br />

nautical mile meter 1,852<br />

statute mile<br />

Mass<br />

meter 1,609.344<br />

To convert from To Multiply by<br />

carat (metric) gram (g) 0.2<br />

grain gram (g) 0.0647989<br />

gram kilogram (kg) 0.001<br />

pound-mass (avoirdupois) (lbm) kilogram (kg) 0.453592<br />

kilogram (kg) pound-mass (avoirdupois) (lbm) 2.20462<br />

long ton kilogram (kg) 1,016.046<br />

short ton ilogram (kg) 907.1847<br />

metric ton kilogram (kg) 1,000<br />

ounce mass (avoirdupois) kilogram (kg) 0.0228349


UNITS OF MEASUREMENT<br />

ounce mass (troy or apothecary) kilogram (kg) 0.0311035<br />

slug<br />

Moment of inertia<br />

kilogram (kg) 14.5939<br />

To convert from To Multiply by<br />

kilogram-meter2 (kg-m2 ) pound-foot2 (lb-ft2 Momentum<br />

) 23.7304<br />

To convert from To Multiply by<br />

kilogram-meter/sec (kg-m/s)<br />

Power<br />

pound-foot/second (lb-ft/sec) 7.23301<br />

To convert from To Multiply by<br />

Btu (thermochemical)/second Watts (W) 1,054.3502<br />

Watt foot-pound-force/second<br />

(ft-lbf/sec)<br />

0.737562<br />

horsepower (hp) foot-pound-force/second<br />

(ft-lbf/sec)<br />

550<br />

horsepower (hp) Watts (W) 745.6998<br />

horsepower (metric)<br />

or chevaux (ch)<br />

Pressure<br />

Watts (W) 735.499<br />

To convert from To Multiply by<br />

atmosphere Newton/meter2 (N/m2 ) 101,325<br />

bar Newton/meter2 (N/m2 ) 100,000<br />

centimeter of mercury (0oC) Newton/meter2 (N/m2 ) 1,333.22<br />

centimeter of water (4oC) Newton/meter2 (N/m2 ) 98.0638<br />

inch of mercury (32oF) Newton/meter2 (N/m2 ) 3,386.389<br />

inch of mercury (60oF) Newton/meter2 (N/m2 ) 3,376.85<br />

inch of water (60oF) Newton/meter2 (N/m2 ) 248.84<br />

lbf/foot2 Newton/meter2 (N/m2 ) 47.8803<br />

lbf/inch2 (psi) Newton/meter2 (N/m2 ) 6,894.757<br />

Pascals (Pa) Newton/meter2 (N/m2 ) 1.0<br />

Newton/meter2 (N/m2 ) lbf/inch2 S<strong>pe</strong>cific heat capacity<br />

(psi) 0.00014504<br />

To convert from To Multiply by<br />

Joule/gram- British thernal unit/pound- 0.238846<br />

degree Celsius (J/goC) degree Fahrenheit<br />

(BTU/lb oF) kilojoule/kilogram- British thernal unit/pound- 0.238846<br />

degree Celsius (kJ/kgoC) degree Fahrenheit<br />

(BTU/lb o S<strong>pe</strong>ed<br />

F)<br />

To convert from To Multiply by<br />

meter/second (m/s) feet/second (ft/sec) 3.28084<br />

foot/second (ft/sec) meter/second (m/s) 0.3048<br />

B.3


B.4 APPENDIX B<br />

Stress<br />

To convert from To Multiply by<br />

Newton/millimeter2 (N/mm2 ) tonf/in2 0.064749<br />

lbf/inch2 (psi) Newton/meter2 (N/m2 ) 6,894.757<br />

ksi<br />

Tem<strong>pe</strong>rature<br />

MPa 6.894757<br />

To convert from To Multiply by<br />

Celsius (T oC) Kelvin (T oK) (T oK) = (T oC) +273.15<br />

Fahrenheit (T oF) Kelvin (T oK) (T oK)= 5/9<br />

(T oF +<br />

459.67)<br />

Fahrenheit (T oF) Celsius (T oC) (T oC)= 5/9 (T<br />

oF – 32)<br />

Rankine (T oR) Kelvin (T oK) (T o Viscosity<br />

K)= 5/9 (T<br />

oR) To convert from To Multiply by<br />

centistokes (cst) meter2 /second (m2 /s) 1 × 10 –6<br />

stoke meter2 /second (m2 /s) 0.0001<br />

foot2 /second (ft2 /s) meter2 /second (m2 /s) 0.092903<br />

centipoise (cP) Newton-second/meter2 (N.s/m2 ) 0.001<br />

centipoise (cP) Pascal-second (Pa·s ) 0.001<br />

poise (P) Newton-second/meter2 (N·s/m2 ) 0.10<br />

poise (P) Pascal-second (Pa·s ) 0.10<br />

slug/foot-second Pascal-second (Pa·s ) 47.88025<br />

lbm/foot-second Pascal-second (Pa·s ) 1.488216<br />

lbf-second/foot2 Volume<br />

Pascal-second (Pa·s ) 4.788<br />

To convert from To Multiply by<br />

barrel (<strong>pe</strong>troleum) US gallon 42<br />

gallon (Im<strong>pe</strong>rial liquid) liter (L) 4.54608<br />

gallon (US liquid) liter (L) 3.78541<br />

liter (L) meter3 (m3 ) 0.001<br />

ounce (US fluid) meter3 (m3 ) 2.957352 ×<br />

10 –5<br />

foot3 meter3 (m3 ) 0.0283168<br />

quart (US liquid) liter (L) 0.946353<br />

Data from:<br />

Mechanical Engineers Reference Book, 11th ed., A. Parrish (Ed.). London: Newnes-<br />

Butterworths, 1978.<br />

Pump Handbook, 2nd ed., J. Karassik, W. H. Ckrutzch, W. H. Fraser, and J. P.Messina<br />

(Eds.). New York: McGraw Hill, 1986.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!