You are on page 1of 192

COLUMN FLOTATION

J. A. FINCH
McGill University, Montreal, Canada

and

G. S. DOBBY
University of Toronto, Ontario, Canada

PERGAMON PRESS
Member of Maxwell Macmillan Pergamon Publishing Corporation
OXFORD • NEW YORK • BEIJING - FRANKFURT
SAO PAULO-'"■ SYDNEY • TOKYO • TORONTO
U.K. Pergamon Press pic. Ilcudington Hill Hall.
Oxford 0X3 OBW. England
U.S.A. Pergamon Press. Inc.. Maxwell House. Fairview Park.
Elmsford. New York 10523. U.S.A.
PEOPLES REPUBLIC Pergamon Press. Room 4037. Oianmen Hotel. Beijing.
OF CHINA People's Republic of China
FEDERAL REPUBLIC Pergamon Press GmbH. Hammerweg 6.
OF GERMANY D-6242 Kronbcrg. Federal Republic of Germany
BRAZIL Pergamon Edilora Ltda. Rua Eca de Oueiros. 346.
CEP 04011. Paraiso. Sao Paulo. Brazil
AUSTRALIA Pergamon Press Australia Pty Ltd.. P.O. Box 544.
Potts Point. N.S.W. 2011. Australia
JAPAN Pergamon Press. 5th Floor. Matsuoka Central Building.
1-7-1 Nishishinjuku. Shinjuku-ku. Tokyo 160. Japan"
CANADA Pergamon Press Canada Ltd.. Suite No. 271.
253 College Street. Toronto. Ontario. Canada M5T IR5

Copyright © 1990 J. A. Finch and G. S. Dobby


All Rights Reserved. No pan of this publication may be
reproduced, stored in a retrieval system or transmitted in
any form or by any means: electronic, electrostatic,
magnetic tape, mechanical, photocopying, recording or
otherwise, without permission in writing from the publisher.
First edition 1990

Library of Congress Cataloging in Publication Data


Finch. J. A. (Jim A.)
Column flotation/.!. A. Finch and G. S. Dobby.—1st ed.
p. cm.

I. Flotation. I. Dobby, G. S. (Glenn Stephen). 1952-


II. Title.
TN523.F46 1990 622\752—dc20 89-48009

British Library Cataloguing in Publication Data


Finch. J. A.
Column flotation.
I. Mining. Fluid power equipment
I. Title II. Dobby. G. S.
622'.028
ISBN 0-08-040186-4

Printed in GreafTiritain by BPCC Wheatons Ltd, Exeter


Contents

PREFACE ix
ACKNOWLEDGEMENTS »

NOMENCLATURE xi»

GLOSSARY xvii
Chapter 1 Background

1.1 Introduction '


1.2 Basic design and operation 1
1.2.1 Design 1
1.2.2 Operation 3
1.3 Introduction to terms 4
1.4 Brief history of the Canadian column 5
1.5 Objective and organization of book 7
1.5.1 Objective 7
1.5.2 Organization 7

Chapter 2 Collection Zone: Gas Holdup, Bubbly Flow and Bubble


Generation

2.1 Gas Holdup 9


2.1.1 Measurement 9
2.1.1.1 Theory of measurement 10
2.1.1.2 Measurement practice 11
2.1.2 Effect of operating variables on gas holdup 13
2.1.2.1 Gas rate and liquid rate 13
2.1.2.2 Frother concentration (bubble size) and solids
concentration 14
2.2 Theory of bubbly flow 16
2.2.1 Slip velocity in a bubble swarm 16
2.2.2 Application of bubbly flow model 17
2.2.2.1 Gas holdup versus gas rate 17
2.2.2.2 Estimating bubble size: d^ versus J 18
2.2.2.3 Effect of frother dosage on bubble size 20
2.2.2.4 Quantifying coalescence 20
2.2.2.5 Limiting conditions: superficial bubble surface rate 22
2.2.2.6 Effect of solids on gas holdup 24
2.2.2.7 Carrying rate 25
2.3 Bubble generation 27
2.3.1 Internal spargers 27
2.3.1.1 Effect of sparger material and surface area: single
spargers 29
2.3.1.2 Multiple spargers 32
2.3.1.3 Frother addition to the gas phase 33
2.3.2 External spargers 34
2.4 Summary of chapter 35

Chapter 3 Collection Zone: The Particle Collection Process


3.1 An expression for the collection rate constant 37
3.2 A collection model 38
3.2.1 Background 38
3.2.2 Particle-bubble collision 39
3.2.2.1 Low particle inertia (Sk = 0) 41
3.2.2.2 Intermediate particle inertia (Sk > 0.1) 42
3.2.3 Particle-bubble attachment 43
3.2.3.1 Particle sliding time 44
3.2.3.2 Attachment efficiency 47
3.2.4 Collection model summary 48
3.3 Applying the collection model: effect of flotation variables 48
3.3.1 Particle size and hydrophobicity 48
3.3.2 Bubble size Z...50
3.3.3 Particle density 52
3.3.4 Viscosity 52
3.3.5 Bubble swarm 52
3.4 Optimum gas rate for collection 55
3.5 The maximum collection rate constant 55
3.6 Summary of chapter 58

Chapter 4 Collection Zone: Mixing

4.1 Mixing models for first-order kinetics 59


4.2 Determination of mixing parameters 61
4.2.1 Axial dispersion coefficient 63
4.2.2 Particle mean residence time 66
4.2.3 Particle vessel dispersion number 67
4.3 Defining collection zone height 68
4.4 Effect of collection zone height to diameter ratio 69
4.5 Summary of chapter 73

Chapter 5 Column Froths

5.1 Froth structure 75


5.2 Gas holdup: effect of variables - gas rate, frother dosage and bias
water rate 78

5.2.1 Effect of gas rate 79


5.2.2 Effect of frother dosage, or bubble size 81
5.2.3 Effect of bias water rate ..83
5.3 Cleaning action and effect of variables 84
5.3.1 Effect of gas rate and bubble size 84
5.3.2 Effect of bias water rate 86
5.3.3 Effect of froth depth 88
5.4 Selectivity 89
5.5 Froth drop back 89
5.6 Carrying capacity 90
5.6.1 Carrying capacity model 91
5.6.2 Experimental carrying capacities 92
5.6.3 Some operating implications of carrying capacity constraints ....97
5.7 Froth stability: effect of reagents and solids 97
5.8 Summary of chapter 100

Chapter 6 Interaction Between Zones


6.1 Column arrangement versus bank arrangement 103
6.2 Defining rate constants 107
6.2.1 Perfect mixing case 107
6.2.2 Plug flow case 108
6.3 Cocurrent column operation 109
6.4 Summary of chapter 110

Chapter 7 Column Scale-up

7.1 Scale-up model 111


7.1.1 Particle collection Ill
7.1.2 Carrying capacity 115
7.1.3 Sparger surface area 116
7.1.4 Conclusions from scale-up example 116
7.2 Laboratory/pilot column testing 116
7.2.1 The pilot column 116
7.2.2 Amenability testing 119
7.2.3 Parameter estimation 123
7.2.3.1 Rate constant 123
7.2.3.2 Carrying capacity 126
7.3 Large pilot column testing 126
7.4 Scale-up case studies 127
7.4.1 Case study 1: bulk Pb/Zn middlings cleaning (Mt. Isa Mines).. 127
7.4.2 Case study 2: Cu-Ni separation (Falconbridge) 129
7.5 Flotation column circuits 130
7.5.1 Selected review 130
7.5.2 Simulation of the effect of some design and operating
variables 131
7.6 Summary of chapter 135
Chapter 8 Measurement and Control
8.1 Measurement 137
8.1.1 Pulp/froth interface position 138
8.1.1.1 Pressure - single measurement 138
8.1.1.2 Multiple pressure sensors 141
8.1.1.3 Float "ZZZ'ZZ'"". 143
8.1.1.4 Temperature 144
8.1.1.5 Electrical conductivity 146
8.1.2 Bias ""!"i48
8.1.2.1 A problem of definition 148
8.1.2.2 Mass balancing for measurement of Jo 149
8.1.2.3 Property profiles for measurement of J 149
8.1.3 Gas rate and gas holdup 151
8.1.3.1 Gas rate 151
8.1.3.2 Gas holdup 151
8.2 Control ' l52
8.2.1 The process matrix 152
8.2.2 Stabilizing control 154
8.2.3 Optimizing control 156
8.2.3.1 Gas rate/gas holdup 156
8.2.3.2 Bias/wash water rate 157
8.2.3.3 Froth depth 158
8.2.4 Circuit control 158
8.3 Summary of chapter 159

APPENDIX A 161

Approximate operating conditions at selected column installations

APPENDIX B 162

Correction for pressure (column height) on gas rate, bubble size and gas holdup

APPENDIX C I63

Example calculations of gas holdup and bubble diameter

REFERENCES 166

INDEX l75

viii
Preface

The authors started work on columns in 1981 when Glenn was awarded a
scholarship from Noranda Mines for a Ph.D. study with Jim on column flotation
scale-up, following the first installation of columns at Les Mines Gaspe. One of the
original patent holders, Mr. Don Wheeler, was instrumental in us getting started. He
showed us, among other things: how to get an interface, the correct design of gas
flowmeter and the importance of a positive bias. By 1983 the work had gained the
financial support of the Natural Sciences and Engineering Research Council of
Canada (Strategic Grants program) and the Department of Energy, Mines and
Resources, Canada. To help stimulate industrial interest a seminar was held to
which Don Wheeler and Dr. Chris Dell (inventor of the "Leeds" column) contrib
uted. As an attendee Mr. Roger Amelunxen showed more than passing interest, as
he was testing a "home made" device back at Gibraltar Mines. The authors collabo
rated with Roger through 1984 and 1985, a time when perhaps the only columns
operating were at Gilbraltar.
In 1985 we were approached by both Cominco Ltd. and Mount Isa Mines
Ltd. to assist in evaluating columns. This offered the first opportunity to test our
ideas on scale-up. At Mount Isa Mines the work was conducted by a graduate
student, Mr. Rodolfo Espinosa-Gomez under the detailed eye of Dr. Bill Johnson.
The work was finished in late 1986 and the first columns were operating at Mount
Isaby late 1987.
Subsequently opportunities were provided by Falconbridge Ltd. (Mr. Doug
Smith and Mr. Burt Huls) and Kidd Creek Mines (Ms. Olga Matwijenko), as well
as further work at Cominco (Mr. Don McKay).
By now Glenn had finished his Ph.D. (1984) and taken a position at the
University of Toronto. At McGill Jim's research was continued jointly with Prof.
Andre Laplante, by Juan Yianatos, Rodolfo Espinosa-Gomez, Cesar Gomez, Rene
del Villar, Manqiu Xu, Shaning Yu and Alejandro Uribe-Salas, while at Toronto a
new group began working under Glenn's direction, including: Ian Flint, Nick
Contini, Maria Falutsu, Steve Wilson and Danny Castillo.
There has been a veritable explosion of activity in the second half of the
1980s to which we are pleased to have contributed. Just in Canada there is expanded
funding and collaborative research among Cominco, Falconbridge (including Kidd
Creek) and INCO, the universities of McGill, Toronto and Laval, and the govern
ment laboratories at CANMET.
The experience of the 80s forms the basis of this book. The authors wish to
thank all those individuals who helped so enthusiastically in plant testing, the
Companies and the granting agencies who have made this endeavour possible.

July, 1989 J. A. Finch G. S. Dobby


Montreal Toronto
r
Acknowledgements

Figures were reproduced, in part or in whole from the following sources:

Canadian Metallurgical Quarterly (publ. Pergamon Press): Figures 2.10; 2.16;


2.17; 2.18; 2.20; 3.12; 3.14; 5.4; 5.5.

Chemical Engineering Science (publ. Pergamon Press): Figures 4.2; 4.3; 4.4;
4.5; 4.6.

CIM Bulletin (publ. Canadian Institute of Mining and Metallurgy): Figures 7.15,
8.19.

Colloid and Surfaces (publ. Elsevier Science Publishers): Figure 2.11.

Column Flotation '88 (Ed. K. Sastry, publ. Soc. Min. Engrs. AIME): Figures
1.3; 4.7; 5.16; 5.18; 7.16.

Computers Applications in the Mineral Industry (Eds. Kostas-Fytas, J.L. Collins


and R.K Singhal, publ. A.A. Balkema): Figures 7.18; 7.20.

CMP Proceedings (publ. Canadian Mineral Processors, division of CIM):


Figure 1.5.

International Journal of Mineral Processing (publ. Elsevier Science Publishers):


Figure 3.3; 3.5; 3.6; 3.8; 3.9; 3.10; 3.11; 5.14.

Journal of Colloidal and Interfacial Science (publ. Academic Press): Figures


2.9; 3.4.

Minerals Engineering (publ. Pergamon Press): Figure 5.20.

Minerals and Metallurgical Processing (publ. Soc. Min. Engrs., AIME):


Figures 4.8; 4.9,4.10, 8.17.

Mineral Processing and Extractive Metallurgy Review (publ. Gordon and


Breach): Figures 5.1; 5.2; 5.3; 5.6; 5.8; 5.9; 5.11; 5.12.

Chemical Process Control An Introduction to Theory and Practice, G. Stepha-


nopoulos (publ. Prentice Hall): Figure 8.1.
K conductivity

K, fraction of bubble surface covered by a monolayer of particles

kc collection zone rate constant

kfc overall (flotation column) rate constant

AL distance between pressure taps on a column

M mass flow

N number of base units in the scale-up calculation (essentially equivalent to


the number of baffled sections in a column)

Nd vessel dispersion number (= E/us|Hc for slurry in columns)

n number of flotation cells in a bank

iIq fraction of colliding particles that collide between 6=0 and 6=9

P pressure

AP pressure difference

Q volume flowrate

R recovery (fractional or percentage)

Rs A/A
<? s

RM total solids mass recovery to overflow, or yield

r radius; radial coordinate

r* dimensionless radial coordinate from bubble centre (radial distance di


vided by bubble radius)

r*-l dimensionless radial coordinate from bubble surface

Reb Reynolds number of a single bubble (equation 2.7b)

Re^ Reynolds number of a bubble in a swarm (equation 2.7a)

Re Reynolds number of a single particle (see equation 4.12)

S length of side of column, if square or rectangular

Sfa bubble superficial surface rate e.g. (cm2 bubble surface/s)/(cm2 column
cross-sectional area)

Sk Stokes number

T temperature

t time
Nomenclature
A column cross-sectional area

A sparger surface area (internal spargers)

C carrying capacity, the maximum measured carrying rate Cf, quoted as mass
of solids recovered per unit time per unit column cross-sectional area e.g.
(g/min)/cm2
C carrying capacity quoted as mass of solids recovered per unit volume of
gas (e.g. g/cm3)
C carrying rate, mass of concentrate solids recovered per unit time per unit
column cross-sectional area e.g. (g/min)/cm2
c. tracer concentration

c particle concentration, number of particles per unit volume of slurry

Dw displacement wash ratio

d. bubble diameter

d particle diameter

dg0 80% passing particle diameter

d column diameter

E axial dispersion coefficient

E. attachment efficiency

Ec collision efficiency

Ec collision efficiency by gravitation

Ec collision efficiency by interception

Eq, Ec calculated assuming Sk = 0 when Sk > 0.05

E., collection efficiency

H collection zone height

Hf froth zone height

Hfc total column height

h height of water level in manometer

Ah difference in water level in manometers (+'ve when upper manometer


water level is higher than that of the lower manometer)

g gravitational acceleration

J superficial velocity (flowrate/column cross-sectional area)

'w Jt-Jf
t. induction time

t time elapsed after tracer impulse injection

t. sliding time (time of particle-bubble contact)

USg slip velocity (relative velocity) between gas and slurry

USp slip velocity (relative velocity) between particle and water

u general term for velocity

ufc terminal rise velocity of a single bubble

u particle velocity

us| interstitial slurry velocity (= SJ(\ - e ))

Vc volume of collection zone

Greek Letters

e gas holdup

e gas holdup in collection zone


e gas holdup in cleaning zone

0 angular coordinate measured from front stagnation point of bubble


8c angle of closest approach for fluid streamlines

6G angle of collision for grazing trajectory

8m maximum angle of contact between sliding particle and bubble, i.e. the
angle at which the particle begins to move radially away from the
bubble surface (if attachment has not occurred)

6n collision angle

9'n 0n that yields ts = t

\Lf water viscosity

Hsl slurry viscosity

p density

pb bubble-particle aggregate density

T mean residence time of collection zone


t nominal mean residence time (= V 7QT = H /J_)

(j) volume fraction solids in a slurry

<}). dimensionless temperature (equation (8.4))

Subscripts (General Usage)


Glossary

Amenability testing: use of a laboratory/pilot column on a given sample for the


purpose of assessing the potential of column flotation.

Bias rate: net flow of water through the cleaning zone (equivalent to difference in
flow of water between underflow and feed) usually expressed as superficial
rate; positive direction is downwards.

Bubbler: equivalent term to sparger.

Bubbly flow: condition characterized by fairly uniform rise of bubbles of fairly


uniform size with little recirculation of liquid or gas; gas holdup generally
below 30-35%; it is the desired flow regime for column operation; see also
chum-turbulent.

Bubbly zone: alternative term for col lection zone; the zone characterized by bubbly
flow.

Carrying capacity: the measured maximum carrying rate, dictated by the froth, for
a given set of conditions.

Carrying rate: mass rate of solids to overflow per unit column cross-sectional area
e.g. (t/hr)/m2.

Churn-turbulent flow: condition characterized by circulation of bubbles and liquid


largely through formation and rapid rise of large bubbles; gas holdup
measurements frequently erratic; it is not the desired flow regime for column
operation; see also bubbly flow.

Cleaning: rejection of solids entrained in water.

Cleaning zone: section of the column above the interface.

Cocurrent: slurry and gas moving in same direction (upwards).

Collection zone: the particle collection section of the column extending from the
interface to the gas sparger (or in the case of laboratory columns, from the feed
port to the gas sparger).

Countercurrent: slurry and gas moving in opposite direction.

Froth zone: alternative term for cleaning zone.

Gas holdup: the volume fraction (or percent) occupied by gas at any given point in
a column; a function of gas rate, slurry rate, bubble diameter, bubble-particle
aggregate density, and solids content of slurry.

Hydrophobic/hydrophilic: general terms used to describe if a particle does/does not


attach to a bubble; degree of hydrophobicity is linked for discussion purposes
with induction time (Chapter 3), contact angle (Chapter 5), and flotation rate
constant (Chapter 6).

XVI
Induction time: time for film between particle and bubble to thin, rupture and form
a 3-phase line of contact.

Laboratory/pilot column: typically 4 to 10 cm diameter by 2 to 10 m high; used in


the laboratory or in-plant for amenability testing and design parameter
determination.

Level: the location of the interface between the collection and cleaning zones; it is
probably better to use froth depth which is unambiguous.

Pilot column, large: typically up to 100 cm diameter by 10 m high; used to confirm


laboratory/pilot column data, to provide operator training, and to evaluate
practical aspects such as sparger life.

Scale-up: selection of size and number of columns for a particular duty, based on
results from laboratory/pilot column or large pilot column tests.

Selectivity: selection of particles depending on differences in hydrophobicity.

Slip velocity: relative velocity between two phases.

Sparger: a bubble generation and distribution system; can be internal (e.g. perforated
rubber) or external (e.g. U.S. Bureau of Mines device).

Superficial rate or velocity: volume flow rate of a species or phase per unit column
cross-sectional area, e.g. (cm7s)/cm2, or cm/s.
Wash water: water added near overflow lip which provides bias water and overflow
water.
Chapter 1
Background

1.1 INTRODUCTION

The flotation column discussed in this book is that patented in the early
1960s by Boutin and Tremblay (Canadian patents 680,576 and 694,547). Early
descriptions of the column and testwork were given by Wheeler (1966) and Boutin"
and Wheeler (1967). This design is sometimes called the 'Canadian' column; given
the number of variations, it could now reasonably be called the 'conventional'
column.
Other designs bearing the name column include: the Leeds column (Dell
and Jenkins, 1976; the patent for this actually pre-dates that of the Canadian
column), the packed column (Yang, 1988), the Flotaire column (Zipperian and
Svensson, 1988), the Hydrochem column (Schneider and Van Weert, 1988), the
column developed by Jameson (1988) and the several designs described in a recent
Russian publication (Tyurnikova and Naumov, 1981).
The emphasis on the Canadian column reflects the fact that there is more
published on these columns and there are more in plant operation than any other
single design.

1.2 BASIC DESIGN AND OPERATION

1.2.1 DESIGN

The column is shown schematically in Figure 1.1 with a photograph of an


installation at Mt. Isa Mines in Figure 1.2. Column is the appropriate description.
Commercial units are typically 9-15 m high and 0.5 to 3.0 m in diameter (the largest
unbaffled column is 2.5 m, the largest baffled column is 3.0 m). The cross-section
may be square (the design supplied by Column Flotation Company of Canada) or
circular (favoured in most 'homemade' units). The side of the square column or
diameter of the circular column is used to designate column size.
Apart from the shape two features which distinguish the column from
other flotation machines are the bubble generation system and the use of wash
water.
Bubble generation is achieved either directly through internal spargers or
after external contacting of gas with water or slurry (referred to here as external
sparging). Internal spargers are made from perforated pipe clad usually in fabric
(e.g. filter cloth) or perforated rubber. An example of an external sparger is that
developed by the U.S. Bureau of Mines (McKay et al., 1988).
Wash water is added into the froth, usually from an array of perforated
pipes located just below the overflow lip.
Wash water

Food
Colloction zone Hc

Gas

Tailings

Figure 1.1 Schematic of flotation column.

Figure 1.2 Installation of three 2.5 m x 15 ra flotation columns for Pb/Zn middlings
cleaning at Mt. Isa Mines (courtesy Mt. Isa Mines Ltd.).
1.2.2 OPERATION

Feed slurry enters about one-third the way down from the top and descends
against a rising swarm of bubbles generated by the sparger. The bubbles collect the
floatable particles, hence the description collection zone (Figure 1.1). This zone is
equivalent to the pulp (or slurry) or recovery zone in a conventional cell. The
collected particles are transferred into a froth stabilized by the wash water. The
wash water's prime role is to clean the froth of particles entrained in the water
crossing with the bubbles from the collection zone. Hence the froth zone is also
called the cleaning zone. (The term washing zone has been used also but it refers
to the zone between the feed and pulp/froth interface (Wheeler, 1983)).
The use of wash water is the key feature which permits high upgrading.
Figure 1.3 illustrates this for coal flotation (Nicoletal., 1988); wash water permitted
the column performance to approach that of the "ideal separation by flotation"
achieved using the release analysis procedure (Dell et al., 1972).
Wash water divides between that going to the overflow (and carrying out
the collected particles) and that going down through the froth to the collection zone.
It is the latter which provides the cleaning action, and the water moving downward
is called the bias water. For efficient cleaning the bias water is normally greater than
zero in the downward direction. Thisistermedoperatingwith'positivebias'. When
the bias is positive the tailings water flowrate is greater than the feed water flowrate
and the bias water supplies the difference so that a pulp/froth interface is maintained.

60

50

~ 40


* 30
u

S 20
Q.

10

5 10 15 20 25
Product ash (%)

Figure 1-3 Effect of wash water in bringing separation close to "ideal separation by
flotation." Wash water rate Jw is volumetric flowrate per unit of column cross-sectional
area, see Section 1.3 (Nicol et al., 1988).
Overall flotation column performance is influenced by interaction be
tween the collection and cleaning zones. An important term in this regard is the froth
dropback, that fraction of the solids entering the froth which is returned to the
collection zone. This improves separation but decreases capacity per unit volume
of the column.

1J INTRODUCTION TO TERMS

Inevitably terms new to the mineral process engineer have accompanied


the introduction of columns. A glossary of terms and a list of symbols are provided.
It is worth briefly introducing some of the terms and symbols here.
In discussing the collection zone gas holdup and superficial gas rate will
be used. Gas holdup is the fraction of gas in a gas-liquid or gas-pulp mixture. The
magnitude of gas holdup is a clue to the hydrodynamic condition of the collection
zone.

Superficial gas rate (or velocity) J is the volumetric flowrate of gas


divided by the column cross-sectional area, or

Ac
Thus the units are, for example, (cm3gas/s)/(cm2 column area), or units of velocity,
cm/s. The superficial velocity is convenient because it can be compared for columns
of different diameter. It will be seen that J is remarkably consistent for different
column sizes. By analogy, superficial rates are introduced for other fluid flows,
giving the general definition

where i may be bias water (B) or wash water (W) or tailing slurry (T), etc.
By combining bubble surface area with superficial gas rate the quantity
superficial bubble surface rate Sb ((cm2 bubble surface/s)/(cm2 column area)) is
derived. This is used to assess solids carrying rate. In the froth zone it is recognized
there is a maximum solids rate, called a carrying capacity C with practical units of
(t solids/hr)/(m2 column), or t/hr/m2.
Some terms are used synonymously, depending on the context. For
example: concentrate and overflow, tailings and underflow; collection, bubbly, and
pulp zone; cleaning and froth zone.
There was a need to decide on units. To try to maximize the readability as
far as the practising engineer is concerned the decision was to use the familiar metric
units rather than to use strictly SI units. Thus the column diameter will be in cm or
m, but bubble size in mm and particle size in \un. On occasion this means there is
a need to convert units but it was felt this difficulty was minor compared to
contemplating a particle size of 2 x 10'5 m attaching to a bubble of 1 x 10'3 m.
1.4 BRIEF HISTORY OF THE CANADIAN COLUMN
Don Wheeler (1983, 1988) has eloquently described the history of the
flotation column. He recalls the frustrating early days solving the technical
problems but not always solving the non-technical ones. Non-technical problems
envelop any new device, particularly such a radically novel design in a field rich
with individual tastes. Don's persistence and his creation of the company Column
Flotation Company of Canada Ltd. kept alive an idea through the lean years of the
1970s.
The surge of commercial interest after about 1981 is illustrated in Figure
1.4 by plotting the number of journal publications per year since 1960. It shows a
typical exponential growth, so that in 1987 some 30 articles were published, while
the average from 1960 to 1980 was less than one a year.
More importantly Figure 1.4 includes some key dates and events. In 1981
the first commercial flotation column in the Western World was installed at Les
Mines Gaspe (Quebec, Canada) by Column Flotation Company of Canada. The
duty was Mo cleaning (Coffin and Miszczak, 1982). The column proved very
effective, a single column stage replacing several stages. The final version of the
circuit in 1987 was two stages (a 0.9m column followed by an 0.45m column)
compared with a circuit which had comprised up to 13 stages of conventional cells.
These columns at Les Mines Gaspe were the source of several studies which
provided basic data on full size units (Dobby, 1984, Yianatos, 1987).
The circuit simplifications and improved metallurgy demonstrated at Les
Mines Gaspe had attracted the Western Canadian Mo producers, notably Gibraltar
Mines. Armed with information from a March 1983 seminar at McGill University
25
First Commercial
20 - Patent Plant tests
Application In West
Opomiska
Les Mines Gaspo
1 C
5
o
10 Column Seminar
3
Cominco, Trail
Q.
d
1960 1970 2000
Figure 1.4 Publication rate and selected events of flotation column history, from 1960
to present.
Roger Amelunxen from Gibraltar Mines was soon realising the potential for
upgrading on a homemade column. Columns rapidly became standard for Mo
cleaning throughout North and South America. Gibraltar Mines extended the test
work to bulk Cu/Mo cleaning. The result was a clear demonstration of improved
separation (Figure 1.5). By 1986, Gibraltar Mines had commissioned a 3 stage
column circuit for bulk Cu/Mo cleaning. This, along with columns in INCO's matte
separation plant (Feeley et al., 1987), Mt. Isa Mines' Pb/Zn middlings cleaning
circuit (Johnson, 1988) and Cominco Polaris' coarse Pb circuit (Kosick et al., 1988),
has led commercial applications away from just Mo cleaning. Worldwide applica
tions by 1988 included at least Cu, Mo, Pb, Zn and Sn cleaning, bulk sulfide
roughing of Au ores, and coal, graphite and phosphate flotation. A partial list of
users is provided in Appendix A.
The interest is continuing to escalate. A seminar with some forty industrial
and research participants was held by Cominco and CANMET at Trail, B.C.,
November 1987 (Feasby, 1987), and 34 papers were given at the International

Range in
Jeed grade

| 99
a
u

f
O
o
oc

96
10 15 20 25 30
Copper in concentrate (

Figure 1.5 Grade-recovery curve for column flotation and mechanical cells at Gibraltar
Mines (Amelunxen, I98S)

Column Flotation Symposium at Phoenix in January 1988 (Sastry, 1988). Ex


amples of improved metallurgy using columns were described at these meetings.
Other attractive features, such as reduced capital and operating costs (due to lack of
mechanical agitation) and adaptability to automatic control, were debated. Specu
lation was heard on such exciting prospects as using nitrogen to control pulp
potential in sulfide flotation; enclosing columns for gas recycle; the judicious use
of some mechanical agitation; and increasing collection rates by using intense
external bubble-slurry mixing.
The flotation column is an example of what research can contribute to the
industry. Until experience has been disseminated, as is largely the case with
operating mechanical flotation cells, the technical literature remains the important
source of information. The authors consider, therefore, that this book on flotation
columns is timely.

1.5 OBJECTIVE AND ORGANIZATION OF BOOK

1.5.1 OBJECTIVE

As in any rapidly evolving subject new terms are introduced, new concepts
developed, or adapted from other disciplines, and theories are created and modified.
It is difficult enough for the researcher to keep abreast but for the plant engineer, who
is the eventual customer for all this effort, the task is daunting.
The intention is to produce a text that can be used by both plant and the R
and D engineers as an introduction to basics, as a reference source and above all a
guide to column testing and practice. (The book does not address column design in
great detail.) To help the reader a summary of each chapter is provided. It is hoped
that the book will provide a common basis and terminology which will help the
technology grow in an orderly way. While theory is included, the practising
engineer is assured that theory is only used to elucidate points of importance in
understanding column operation: for example, the interaction of gas rate and bubble
size in the performance of the collection and cleaning zones.

1.5.2 ORGANIZATION

The book is organised into chapters. Chapters 2-6 examine the fundamen
tals, linking them to industrial experience, while Chapters 7 and 8 consider aspects
of pilot-scale testing, scale-up and control.
Chapters 2 - 4 are devoted to the collection zone. In Chapter 2 the basics
are covered. Extensive use is made of gas holdup and superficial gas rate. A model
is developed to define the bubbly flow regime which is the desired regime for this
zone. The model is used to estimate bubble diameter and limits to solids removal
rate (or carrying capacity). Bubble generation is also discussed.
Chapter 3 examines the collection process. A model is developed and used
to assess the effect of variables (e.g. bubble diameter, particle diameter, gas rate) on
flotation rates.
Chapter 4 examines mixing characteristics of the collection zone. It is
shown that the degree of mixing can be estimated from column dimensions. This
introduces the possibility of scaling up columns and of examining the effect of
column geometry. For example, arguments are advanced showing that a reasonable
height-to-diameter ratio of the collection zone is in the order of 10:1.
Chapter 5 is devoted to the froth (or cleaning) zone. The emphasis is on a
qualitative description of froth phenomena and their impact on cleaning. Recom
mendations for operating conditions to maximize cleaning are presented and
speculation on the possible selectivity of column froths is offered. Dropback of
solids from the froth and the carrying capacity limitation of froths are introduced.
Interaction between the collection and cleaning zones is described in
Chapter 6. Arguments are advanced for increased selectivity in columns compared
with mechanical cells. The relationship between collection zone rate constant and
overall rate constant is examined, and the impact of dropback on metallurgy and
capacity assessed.
Chapter 7 addresses scale-up. The procedure is described which combines
knowledge of mixing, dropback and carrying capacity with parameters determined
from pilot scale tests. Pilot scale testing is described, and the scale-up approach is
illustrated with examples from Mt. Isa Mines and Falconbridge Limited. A selected
review of column circuits is presented.
Chapter 8 evaluates the current state of column instrumentation and
control and outlines future possibilities. Emphasis is on control of level and bias
which pose novel problems in flotation columns.
Chapter 2
Collection zone: Gas holdup, bubbly flow
and bubble generation

2.1 GAS HOLDUP

When gas is introduced into a column liquid (or slurry) is displaced. The
volumetric fraction displaced is called the gas holdup 6g. The complement (1 - 6g)
is the liquid (or slurry) holdup.

2.1.1 MEASUREMENT

Gas holdup can be measured in a number of ways (see Figure 2.1). Method
(a) measures the gas holdup for the whole vessel ('overall' holdup) while methods
(b) and (c) measure holdup over a given section of the vessel ('local' holdup). In
method (b) the given section is defined by the distance between the pressure tapping
points; in method (c) it is defined by the signal path between the probes. Methods
(b) and (c) can be used to measure axial variation in gas holdup. Method (c) can be
modified to measure gas holdup at specific points within the column.
Method (b), the pressure method, is selected for examination.
This method is the simplest; method (a), forexample, is impractical with a froth, and
method (c) needs calibration, usually against gas holdup determined from pressure
methods.

GAS HOLDUP MEASUREMENT

□ D
AP

l
tJt-t .1 tit
Gas

a) Level rise b) Pressure c) Sensor


difference (e.g. absorption
of X-rays,
conductivity)

Figure 2.1 Methods of measuring gas holdup.


2.1.1.1 Theory of measurement

The practical case of gas-slurry will be considered. To make the solution


tractable it is assumed that the dynamic component of the pressure is negligible and
that the bubbles are lightly loaded, and consequently the bubble-particle aggregate
density is negligible. The pressure above atmospheric at A and B (Figure 2.2a) is
then given by

Pa =
and
PB =
where ps, is the slurry density, and € , egB are the gas holdup above A and B
respectively (the product L(l - €g) is the equivalent height of slurry without gas).
Therefore, the pressure difference between A and B, AP, is

AP = ps]gAL(l-eg)

where €g is the gas holdup between A and B. Upon rearranging, e is given by

AP
6=1- (2.1)

_ liquid level

Ah

h,

AL

J A

Gas

Figure 22 Measurement of gas holdup by pressure difference: (a) general; (b)


using water manometers.

10
Note that € is measured locally over distance AL and that gas holdup in other parts
of the column is not a factor. By repeating the measure at intervals along a column,
the profile of € with height can be established.
If water-filled manometers are used to measure pressure (Figure 2.2b),

where Ah is positive when the upper manometer level is higher than the lower
manometer level. (When Ah is positive the bulk density of the collection zone is less
than that of water, when Ah is negative the bulk density is greater than that of water.)
Then e is given by
s

p AL

2.1.1.2 Measurement practice

In the three phase system it is necessary to know the slurry density between
the pressure taps; the sensitivity of € to ps] will be illustrated.
Consider the case where pressure is measured with water manometers, and
it is assumed p = pw. Equation (2.2) then reduces to

03)
8 AL

This is the solution for a gas-water system. With slurry, equation (2.3) gives an
apparent holdup e b . The actual holdup eg is related to €g|app by

(2.4)
. e )
g

The relationship is explored in Figure 2.3. There is a significant difference between


the actual and apparent holdup. For example, consider ps| = 1.3 g/cm3 (equivalent
to 30 % solids by weight for chalcopyrite) where an actual gas holdup of 20% will
be read as an apparent holdup of -4%. The same problem arises if pressure
transducers are used: by inspection of equation (2.1) if the calibration assumes ps|
= p then, for the above example, the apparent holdup is calculated to be -4%.

11
35

30

Co 25
o
J

Q. c 20

o
15

10

-10 -5 10 15 20
Apparent holdup, € g I. (%)
|App

Figure 2.3 Actual gas holdup versus apparent gas holdup measured when using
water manometers.

The example used was extreme (assuming p . = p ) but illustrates the


point. It is necessary to know ps] if accurate gas holdup values are required.
In flotation the problem is further compounded since solids partition
between the water and the bubbles. This can introduce an axial variation in p ., and
the assumption of a zero density for the bubble-particle aggregate may be invalid
in some cases. A possible solution in the plant is to measure AP over a section near
die bottom of the column, where the bubbles are likely to be lightly loaded, and use
the tailings slurry density to estimate the slurry density within that section. If the
solids are coarser than about 30 |im (d^) then corrections for relative downward
velocities of the particles and the liquid must be made (Dobby et al., 1988). The
problem of allowing for bubble-particle aggregate density is addressed later
(Section 2.2.2.6).
Manometric methods have some practical problems. Bubbles and solid
particles can enter the manometer arm if precautions are not taken. In a gas-liquid
system bubbles can be excluded either by a porous plug (e.g. steel wool) or by
connecting the arm at a small downward angle to the vessel. In a gas-slurry system
solids can be eliminated by maintaining a small downward flow of water. In
laboratory/pilot columns care is required to ensure that this flow of water through

12
the manometers does not represent a significant proportion of the total water
entering the column.
Pressure transducers mounted on a probe or on the side of the column
eliminate most of the operational problems of water manometers, as well as
providing a continuous analog signal that can be used for monitoring and/or control.

2.1.2 EFFECT OF OPERATING VARIABLES ON GAS HOLDUP

2.1.2.1 Gas rate and liquid rate

A useful method of defining gas and liquid rate is the volumetric flow rate
per unit cross-section, or superficial velocity, as indicated in Chapter 1. The
superficial velocity is useful since it can be compared for columns of different
cross-section.
Volumetric gas flow varies with static head. In a column of height 12 m
pressure at the gas inlet is about twice that at the overflow. Unless noted otherwise,
values of gas rate quoted hereon refer to the column overflow, i.e. at atmospheric
pressure (see Appendix B).
The relationship between e andJ is used to define the flow regime (Shah
et al., 1982). Figure 2.4 shows the general relationship. Gas holdup increases
approximately linearly then deviates above a certain range of Jg. The linear section

09
CO

(A
(0
o

bubbly flow chum - turbulent


regime flow regime

Superficial gas velocity, Jg (cm/s)

Figure 2.4 Gas holdup as a function of gas rate, general relationship.

13
is characterized by a homogeneous distribution of bubbles of fairly uniform size
rising at a fairly uniform rate. This is called the bubbly flow regime. Above the
transition J^ gas holdup becomes unstable and the flow is characterized by large
bubbles rising rapidly, displacing water and small bubbles downward. This is the
churn-turbulent flow regime. (In small diameter columns, dc < 0.1m, the large
bubbles may fill the cross-section giving a slug flow regime.) Data on columns up
to dc = 2.5 m confirm the relationship in Figure 2.4.
The desirable column operation is in the bubbly flow regime. This places
an upper limit on gas rate, as will be discussed later.
Figure 2.5 shows the result of increasing J, countercurrent to the bubbles
(the usual column operation). As J, increases for a given J , € increases. This is
expected since bubble rise velocity (relative to a stationary observer) is decreased.
Increasing J, will decrease the maximum J that can be tolerated for operation to
remain in the desired bubbly flow regime.

2.1.2.2 Frother concentration (bubble size) and solids concentration

The effect of gas and liquid rate was illustrated (Figure 2.5) at a fixed
frother concentration. The addition of frother up to a certain concentration has a
pronounced impact on reducing bubble size. A reduced bubble size means reduced
bubble rise velocity and consequently gas holdup is increased.
The impact of frother concentration is illustrated in Figure 2.6, where gas
holdup, at a given J , increases significantly from 0 to 15 ppm Dowfroth 250C. The
increase in gas holdup levels off above a certain frother dosage (usually about 30
ppm), corresponding to the known trend of bubble size with frother concentration
(Klassen and Mokrousov, 1963); this is addressed further in Section 2.2.2.3.
The role of solids is more difficult to describe. A general consideration is
that solid particles may promote or retard coalescence depending upon their surface
properties. Specific to flotation is that the solids distribute between the liquid phase
(i.e. remain in suspension) and the gas phase (i.e. attach to the bubble).
Mineral solids in suspension give a slurry density ps( and a slurry viscosity
p.sl which have opposing effects on the bubble rise velocity. Solids attached to the
bubble give a bubble-particle aggregate density pb greater than zero, causing
bubble rise velocity to decrease and, therefore, gas holdup to increase. Predicting
the holdup in the presence of solids is therefore complicated.
The discussion so far has given a qualitative guide to the effect of variables
upon gas holdup. A theoretical section follows which tries to quantify the effects.

14
0.5 1.0 1.5
Superficial gas velocity, Jg(cm/s)

Figure 2£ Gas holdup versus gas rate, effect of downward liquid velocity. 15 ppm
Dowfiroth 250C.

40

12 3 4
Superficial gas rate, 4g (cm/s)

Figure 2.6 Gas holdup versus gas rate, effect of frother (Dowfiroth 250C) dosage. J(
=0.5 cm/s.

15
2.2 THEORY OF BUBBLY FLOW

2.2.1 SLIP VELOCITY IN A BUBBLE SWARM

The slip velocity Us is the velocity of one phase relative to another. In this
case, USg is the velocity of the gas phase relative to the liquid (or slurry) phase, and
is defined as

The +/- sign refers to countercurrent flow/cocurrent flow respectively. (The


convention here will be gas is positive upwards, liquid positive downwards, since
this is the common usage.)
The slip velocity is related to the system variables. For bubble sizes of
interest here, d^ 2 mm (Reb < 500), a suitable expression is an adaptation (Yianatos
et al., 1988) of the multi-species hindered settling equation of Masliyah (1979),
written below for the gas-slurry system

v _gdb2(Ps|- Pb> F(1-£E)


O 1 O ft I / t m f\ 9 C V\

where

Rebs=—§E-5! *■ (2.7a)

and
F(l-e) = (l-e)ml (2.8)

and (Richardson and Zaki, 1954)

m = [4.45 + 18(d,/dc)] Re^'


-0.1
1< Reb< 200 (2.9a)

•O.I
m = 4.45 Re^' 200< Reb< 500 (2.9b)

and

where Ub is terminal rise velocity of a single bubble (this is calculated using equation
(2.6) and assuming € =0).
Equations (2.5) and (2.6) are a model of bubbly flow. The model offers a

16
means of extracting difficult to obtain information, e.g. bubble size and effect of
solids, from readily measured data, €g, Jg and Js, (or J,).
2.2.2 APPLICATION OF BUBBLY FLOW MODEL
2.2.2.1 Gas holdup versus gas rate
Consider the two phase gas-liquid system, where equation (2.6) reduces to
(2.10)
0.687v
18(1,(1+0.15 ReJ00')
If o^ is known, Us can be calculated for a given €g. (This requires an iterative
solution because rI^ = f(U ) and Reb = f(Ub)). At a certain 6g, USg from equation
(2.10) equals Us from equation (2.5), for a given Jg, J,. This is the predicted € for
the given conditions. By repeating at various Jg a relationship between €g and Jg is
predicted.
The method of solution is illustrated graphically in Figure 2.7. The USg
versus e from equation (2.5) is plotted for J = 1.2 cm/s and J{ = 0.5 cm/s (curve 1).
20 40 60 80
Gas holdup, £9(%)
Figure 2.7 Illustration of solution for bubble size. 'Definition* is equation (2.5),
'Settling Equation' is equation (2.10).
17
The USg versus |lg from equation (2.10) is plotted for various (^(curves 2,3 and 4).
The solution for eg is where equation (2.5) intercepts equation (2.10). (There are two
solutions; the one at the lower e is for conditions in the collection zone and the one
at the higher € is for conditions in the cleaning zone.) Thus, as an example, for d
= 1 mm, € - 12%.
Provided d,, is known the model provides a means of predicting € versus
J . Figure 2.8a shows the predicted relationship corresponding to theEdata in
Figure 2.6 where db was measured (photographically) to be 1 mm at J = 1 cm/s and
15 ppm frother. Clearly the predicted 6g does not follow the approximately linear
increase with J but instead increases more rapidly. This is because d. increases with
Jg. Figure 2.1$b shows the much improved fit by introducing the empirical
relationship

db = CJgn (2.11)
where n = 0.25 has been found suitable for the systems studied thus far (Dobby and
Finch, 1986c); see also Section 2.3.1.1.
In Figure 2.8b the end-point of the predicted e versus J line represents the
maximum gas holdup €gmax which can be tolerated for flow to remain in the bubbly
regime. For 6 > eg max equations (2.5) and (2.10) do not intercept and no solution
can be found. This e max is similar to the "flooding point" derived from drift-flux
analysis (Wallis, 1969). This maximum places important restrictions on the
allowed gas rate, bubble size and liquid rate combinations with ramifications which
are explored later.

2.2.2.2 Estimating bubble size: d. versus J

In order to predict € versus Jg, db was required. If e and J are known,


dfe can be estimated. The technique is to repeatedly substitute estimates of d in
equation (2.10) until the predicted Us from equation (2.10) equals the measured
USg from equation (2.5). This can be illustrated using Figure 2.7 again. Suppose
€g were measured at 12%, then for these conditions the estimated db is 1 mm.
Figure 2.9 shows a test of this method of estimating d^ Mean bubble size
(the Sauter mean) was measured photographically in a variety of laboratory
columns over a range of conditions and compared with the estimated d (Yianatos
etal., 1988c). The estimated c^ is within ±15% of the measured db. This technique
has also been applied to data from mechanical cells (Manqiu and Finch, 1988a).
The success of this method of estimating o^ alleviates the tedium and
difficulty of photographically measuring d^ Also, it offers a means of estimating
d)j in opaque columns, and can be extended to gas-slurry systems (Dobby et al.,
1988).
Using this method of estimating bubble size the db versus J relationship,
equation (2.11), was established from laboratory scale tests. Preliminary evidence
from large columns confirms equation (2.11).
In Appendix C the step-by-step calculations for estimating bubble diame-

18
a.
3

a
a
a

1 2 3 4
Superficial gas velocity, J, (cm/s)

Figure 2.8a Predicted gas holdup versus gas rate relationship (data from Figure 2.6):
djj is constant (db = 1.0 mm).

12 3 4
Superficial gas velocity, J9(cm/s)

Figure 2.8b As in 'a' but db= CJ 025 (db=1.0mm at J =lcm/s).

19
.•'-15%

0.5 1.0 1.5 2.0

Bubble diameter, predicted (mm)

Figure 2.9 Comparison of estimated bubble diameter and photographically measured


bubble diameter (Yianatos et al., 1988c).

ter in gas-water and gas-slurry systems are described.

2.2.2.3 Effect of frother dosage on bubble size

The bubble size estimation method can be used to assess the effect of
frother dosage upon bubble diameter in a sparged column. An example of this is
given in Figure 2.10, which shows the results of bubble size measurements in
solutions of water-frother for two sparger types: filter cloth and porous glass frit.
There is clearly a strong effect of frother dosage at levels below about 20 ppm, which
is similar to the trend reported by Klassen and Mokrousov (1963) for bubbles
issuing from a single orifice.

2.2.2.4 Quantifying coalescence

Another application of this technique of bubble size estimation is to study


coalescence. Some flotation reagents may modify the bubble surface and promote
coalescence, which is undesirable as it reduces the specific bubble surface area.
Coalescence can be quantified as a change in gas holdup Ae , or change in bubble
size Adjj.
An example of a class of collectors which can promote coalescence is the

20
0 10 20 30 40
Frother dosage (ppm)
Figure 2.10 Illustration of the effect of frother (Dowfroth 250C) dosage upon d^ J =
1.3 cm/s (Flint etal., 1988).
• Nlobee recycle water
o Montreal tap water
10 20 30
Fatty acid (ppm)
Figure 2.11 Change in gas holdup upon adding fatty acid to 20 ppm frother (TEB)
(Espinosa et al., 1988d). Two water types are illustrated. Niobec is a Nb-mineral
processing plant, near Chicoutimi, Quebec.
21
fatty acids. Figure 2.11 shows the result of adding fatty acid (Actinol FA2 in a 3:1
ratio with lannagol emulsifier) to a solution of 20 ppm frother. The gas holdup
relative to that at zero fatty acid decreased by about 50% at fatty acid concentrations
20 ppm and greater. For these conditions a 50% decrease in e corresponds to a 70%
increase in o^ or nearly 500% increase in bubble volume. 8

2.2.2.5 Limiting conditions: superficial bubble surface rate

In the discussion of Figure 2.8b the concept of a maximum gas holdup for
operation in bubbly flow was introduced. Returning again to Figure 2.7 the impact
of this on imposing a minimum bubble size at a given gas rate, or conversely, a
maximum gas rate at a given bubble size can be illustrated. In Figure 2.7 the U
versus €g from equation (2.10) for db = 0.6 mm has only one intercept (tangent) to
the U versus €g from equation (2.5), at an € of about 30%. Any db< 0.6 mm will
not yield a solution (equations (2.5) and (2. l6) will not intercept). Therefore, (^ =
0.6 mm is the minimum bubble diameter dfc min at J, = 0.5 cm/s and J = 1.2 cm/s; or
conversely, J = 1.2 cm/s is the maximum gas rate if db =8 0.6 mm and
J,=0.5 cm/s. Figure 2.12 shows the calculated boundary conditions for a gas-water
system.

Figure 2.12 shows that at a given J., there is a restriction on the allowed J ,
1 g

O.S 1.0 1.5 2.0

Superficial liquid velocity, JL(cm/s)

Figure 2.12 Gas rate, liquid rate, bubble size limiting combinations for a gas-water
system.

22
db combinations. For example, if J, = 1.5 cm/s and c^ = 1 mm then the maximum Jg
(J ) is about 2.1 cm/s. If d. is smaller, J gmax
_ must decrease. This has two
g man'
important consequences: (a) the surface area rate of bubbles, which controls solids
removal rate (or carrying rate), may not increase with decreasing db as is often
implied, and (b) since the collection rate constant is proportional to J but inversely
proportional to db (see Section 3.1), decreasing c^ may not yield higner collection
rates. Point (b) is discussed in Section 3.5. To address point (a) superficial bubble
surface rate Sb ((cm2 bubble area/s)/(cm2 column area)) is introduced. Sb is given
by gas flow rate multiplied by the ratio of bubble surface area to bubble volume all
divided by column cross-sectional area. This reduces to
Figure 2.13 shows Sb versus bubble size for various gas rates. The limit
imposed by the bubbly flow constraint is the a-b line. The a-b line is given by
6J
(2.13a)
b max
300
&ro
"2.5
a-b: Limitation for operation
9 In bubbly flow regime
o
to
\2.0
200
(0 ^J
0) »
(0 J9
100
a.
0.6 0.8 1.0 1.2
Bubble diameter, db(mm)
Figure 2.13 Bubble surface rate as a function of bubble size; line a-b represents maxi
mum bubble surface rate Sb max (e.g. given by maximum gas rate at the given bubble
size).
23
(see Figure 2.12), or conversely

^= (2.13b)
The maximum Sb decreases with db (at least for a^ < 1.5 mm). Bubble surface rate
has a direct impact on carrying rate, which is explored in Section 2.2.2.7.
Figure 2.12 also shows there is a maximum J for a given d, J combina
tion. To illustrate, if db = 0.8 mm and J = I cm/s the maximum J, is about 2.5 cm/
s. This restriction on J, has an impact on system capacity. For example, if a material
has a high collection rate constant the high capacity this apparently implies may not
be realisable because of the restriction on the downward slurry velocity. A possi
bility in this situation is to use cocurrent gas-slurry flow, which is addressed in
Section 6.3.

2.2.2.6 Effect of solids on gas holdup

Solid particles in flotation distribute between the liquid phase and the gas
phase. Slurry density and viscosity often have approximately equal and opposing
effects on bubble rise velocity, and therefore on e (the exception would be in the
case of fine clay-type particles where viscosity effects may dominate). Therefore,
the attention here is focussed on bubble-particle aggregate density.
An increasing bubble-particle aggregate density pb will decrease Us and
thus increase € . To estimate p., assume that: particles are small relative to the
bubble, each particle occupies d'2 of bubble surface (Szatkowski and Freyberger,
1985), and the bubble loads to a fraction K of a monolayer. Then

K. 71 d p
p = _! LIE. (2.14a)
b (L+K.Ttd
D 1 p

For db » K(jid , which is usually the case, equation (2.14a) reduces to

71 d p
pb = KI__OL (2.14b)
^)

which is equivalent to mass of solids carried per unit volume of gas. According to
Jameson (1986) the maximum loading corresponds to Kl = 0.5 (or 50%) giving a
maximum bubble-particle aggregate density p. of

71 d p
PblM*= - -=-=■ (2-15)
2 <K
Table 2.1 shows the effect of pb max on 6 for db = 1 to 2 mm. At d^ = 1 mm, e

24
increases by more than 50%. The increase becomes less significant with finer par
ticles. Nevertheless, bubble loading will need to be considered in bubble size
estimation, and the increase in € may have to be considered in slurry residence time
calculations.
The increase in gas holdup with bubble loading may partially account for

Table 2.1 Effect of Bubble Loading on Gas Holdup

(Adapted from Yianatos et al., 1988c)

Conditions: Jg=1.0cm/s, Js,= 1.0cm/s,psl= 1.19g/cm3.Hsl = 0.0113 g/cm-s,and


for equation (2.15), d = 50 ^m, p = 5.0 g/cm3

the observation that increases in feed pulp density cause large bubbles to form and
break through the froth (so-called 'burping') (Hoffert, 1987; Kosick et al., 1988).
This observation suggests transition to churn-turbulent flow. In such cases, the
increase in feed solids rate may have increased bubble loading to the point where
the associated increase in gas holdup exceeded that compatible with bubbly flow.
Another contributing factor may be increased viscosity of pulp, especially near the
inlet. In laboratory/pilot columns burping with high percent solids also occurs and
is sometimes associated with a second interface appearing well below the feed inlet
(del Villar, 1987). Decreasing gas rate cures the burping, and the second interface
tends to migrate up towards the primary interface. The fact that a decreased gas rate
eliminates the problem suggests its origin is the transition from bubbly to churn-
turbulent flow.

2.2.2.7 Carrying rate

At a given superficial gas rate and bubble loading there is a certain mass
rate of solids that can be carried. A convenient unit is the mass of solids carried per
unit time per unit column cross-sectional area, or carrying rate Cr, given by

K d D J
Cr= K1_11P_I (2.16)

25
In principle the system can be adjusted to yield a maximum C, namely
(2.17a)
or
(2.17b)
(Later, as a practical measurement of the maximum carrying rate, the symbol C , or
carrying capacity, is introduced; C is not necessarily equal to C , which isW
theoretical maximum rate.)
Strictly, loaded bubbles will exhibit a larger dbmin compared with unloaded
bubbles, which can be allowed for in the bubbly flow model. The increase is usually
small and the effect will not be pursued here; for details see Manqtu et al. (1987).
Figure 2.14 gives model results for C when d = 2 to 25 \im and p =
7.5 g/cm3 (e.g. galena). For the finest size the largest C occurs between bubble
sizes 1 to 1.5 mm (corresponding to the range of maximum Sb, Figure 2.13). As
particle size increases the optimum bubble size increases. It is assumed that K = 0.5
and does not change with bubble size, both of which are questionable.
Direct experimental verification of Figure 2.14 will prove elusive since it
is virtually impossible to maintain a system at its dbmin, i.e. on the verge of transition
to churn-turbulent flow. Nevertheless, the indications are that going to fine bubbles
(db« 1 mm) does not offer a means of increasing C
0.5 1.0 1.5
Bubble diameter, db (mm)
Figure 2.14 Maximum collection zone carrying rate given by equation (2.17) for
conditions desribed in text.
26
2.3 BUBBLE GENERATION

The method of bubble generation is another feature which distinguishes


flotation columns from conventional machines. The most common method of
generation is through an internal sparger (or bubbler) near the base of the column.
A second method of bubble generation is that where gas and liquid (or slurry) are
brought into contact external to the column. Discussion will be divided between
internal and external spargers.

2.3.1 INTERNAL SPARGERS

There are two broad categories of internal sparger: porous spargers (such
as sintered glass) and single and multinozzle spargers (such as punched plate).
Figure 2.15 shows a summary of data for a variety of systems (see Table 2.2). All
the porous spargers exhibit a similar gas holdup which is higher (i.e. bubbles are
smaller) than with multinozzle spargers. The results are for water and the bubble
diameter with the porous spargers is about 2 mm (at J =1.5 cm/s).
Porous spargers are used in flotation columns. Figure 2.15 implies there
is no major difference between rigid or flexible (perforated rubber) spargers. In
practice, with slurries and plant waters flexible spargers are preferred as they are
less prone to plugging by particles or sail deposits. The common flexible spargers

30

X «

P 0

a.

520
Pofoui ga» *

a
a
a

and ntutttnoiil*

10
ISftait *t «i.«

246
Superficial gas velocity, Js(cm/s)

Figure 2.15 Gas holdup in bubble columns with porous and multinozzle gas spargers
(air, N, - water system). See also Table 2.2.

27
Table 2.2
Columns, gas distributors and experimental conditions for Figure 2.15

28
Figure 2.16 Scanning electron micrographs of two porous sparger materials, (a)
perforated rubber: hole diameter = 80 \lm, density = 40 holes/cm2; (b) filter cloth: hole
diameter and density indeterminate (Manqiu and Finch, 1988b).

are made from rubber or fabric, for example a neoprene cloth. The material is
wrapped and secured around perforated steel pipe, the resulting spargers being
arranged in horizontal or vertical arrays.

2.3.1.1 Effect of sparger material and surface area: single spargers

The sparger material (its flexibility, hole size, shape, density of holes (no.
of holes/unit area)) and sparger surface area can be expected to affect performance
(e.g. bubble size).
Results of systematic tests are available on three materials: stainless steel,
filter cloth and rubber (some material texture characteristics are given in Figure
2.16). Sparger surface area (A ) was varied from 20 to 215 cm2 and column cross-
sectional area (A ) from 7.5 cm2 to 78.5 cm2 (dc = 3.08 to 10 cm). Frother
concentration was 15 ppm Dowfroth 250C. It was determined that gas holdup
decreased as the ratio Rs (= AJ\) increased (Figure 2.17), i.e. bubble diameter
increased as R increased, for both rigid (Figure 2.17a) and flexible (Figure 2.17b)
spargers. At R > 1, there was a tendency to produce a less uniform bubble size and
6 versus J started to deviate from linear. This is evidence of transition from bubbly
to churn-turbulent flow.
By observation, not all the sparger surface is active at the same time. For
example, with vertical spargers initially gas emerges from near the top, and as gas
rate is increased the bubble producing surface expands downwards. This poses
some difficulty in defining an active as opposed to the total surface area A . Despite
this complication, bubble diameter (estimated as described in Section 2.2.2.2) was
found to depend on RsJ (Figure 2.18), following

(Rs < 1) (2.18)

where n ~ 0.25 (see dashed line).


Equation (2.18) is a generalization of equation (2.11) which applies to a

29
30

25

20

15

10
o

0 O5 1.0 1.5 2.0


Superficial gas velocity, Jg(cm/s)

Figure 2.17a Gas holdup versus gas rate as a function of R (Manqiu and Finch, 1989
(a) stainless steel: hole diameter » 50fim, density = 10 holes/cm2;

30

25

20

a.
3 15

CD
10
ID
0

0 0.5 1.0 1.5 2.0

Superficial gas velocity, Ja(cm/s)

Figure 2.17 b As in (a), but using filter cloth.

30
1.5

I 10

3
CO

O.S
0 0.5 1.0 1.5 2.0

Gas flowrate per unit area of sparger, Rs-Jg (cm/s)

Figure 2.18 Bubble diameter as a function of (Rs J ); (^ calculated by method


described in Section 2.2.2.2 (Manqiu and Finch, 1989).

constant R . The constant C, will depend, as with C in equation 2.11, on frother


concentration principally. (In Figure 2.18 C, = 1 was used to generate the dashed
line).

The product (Rs J ) can be expanded

s 8 As Ac As

i.e. R J is equivalent to volumetric flowrate of gas per unit surface area of sparger.
The conclusion that bubble diameter depends on volumetric flowrate of gas per unit
surface area of sparger was also reached by Clingan and MacGregor (1987)
although the basis for their conclusion was not given.
Sparger material has a modest effect on bubble diameter and filter cloth
tends to give the largest bubbles (Figure 2.18). Clingan and MacGregor (1987)
tested rubber, ceramic and fabric and commented only on the effects of As, not
material type (and in fact selected fabric). While some spargers will produce
smaller bubbles than others, it should be remembered that metallurgical perform
ance is controlled by both bubble size and gas rate, and that generation of larger
bubbles can often be compensated for by operation at higher gas rates. Higher gas
consumption will obviously be of concern when gas other than air is being utilized.
An implication of the sparger tests to date is that hole size and density over
a broad range (compare filter cloth with rubber, Figure 2.16) is not important. In
industrial columns hole densities from 4 to 40 holes/cm2 are employed (Feasby,

31
1987) with no obvious effect being reported (40 holes/cm2 is about the upper limit
to prevent tearing). A limited effect of hole density may be because several holes
contribute to each bubble. The bubbles (o^ ~ 1 mm) typically are 20 times the size
of the holes (hole ~ 50 nm) and thus an overlap between holes in generating bubbles
can be expected.
It must be stressed that a systematic study of hole density on a single
material needs to be conducted to evaluate this parameter. Recent studies have
shown a significant dependence of the e versus J relationship on the permeability
of filter cloths (Yianatos, 1988). % e
In terms of their role in controlling o^ (at a given J ) the most important
factor is frother concentration (at least up to 15-30 ppm for most frothers), second
is Rs and third is sparger material. This has implications in scale-up. Much pilot-
scale testing is performed with stainless steel spargers; Figure 2.18 shows this
should not pose any difficulties in creating similar db with subsequent commercial
use of rubber or fabric. A criterion for scale-up is preserving similar d., J since this
maintains the surface area rate (equation (2.12)) and collection rati constant
(equation (3.2)). To preserve db, J means preserving Rs, i.e. the sparger surface area
should increase in proportion with the column cross-sectional area. This rule is the
intuitive one. A survey of users (Feasby, 1987) reported volumetric gas rates per
unitareaofspargerbetween0.13and0.31 L/min/cm2(atSTP),orRsJ between2.2
and 5.1 cm/s. Allowing for pressure in the full size units, the range It the sparger
(which is the correct value for comparison) is about 1.1 to 2.5 cm/s, essentially the
same as in Figure 2.18. The predicted bubble diameter, knowing only this R J , is
1 to 1.3 mm, which is about the expected value. Since usually sufficient frother is
present to give close to the minimum db, estimation of dfa from Figure 2.18 may be
a reasonable initial approach in many situations.
While sparger material may not be vitally important for d^ it is a vital factor
in sparger life in the industrial environment of abrasive slurries and corrosive
waters; sparger life of at least 2 months is probably required. Flexible spargers are
invariably used. These act to prevent build-up of scale and to prevent lodgement of
particles in the holes, both of which rapidly occur with rigid spargers. Choice of
material in this respect is one of extensive trial-and-error. Some operators use a
double layer cloth sparger, Gibraltar Mines (Amelunxen, 1985) used a strong nylon
layer, to absorb sudden bursts of air, covered with a nylon-canvas cloth. Suttili
(1987) has described a system for changing spargers on-line.
The data analysed thus far is for a single sparger. The arrangement of
spargers may be a factor.

2.3.1.2 Multiple spargers

Full-size columns will need multiple spargers to achieve the required


surface area. It is not clear whether the arrangement and packing density of spargers
is important. Data from systematic evaluations are limited. Figure 2.19 shows the
arrangement of three ceramic ball spargers in a 10 cm diameter column (Manqiu,

32
20

2 10
o

Jt .0l3 cm/s
frolhor. 15 ppm

"0 0.5 10 1.5 2.0 2.5


Superficial gas velocity, Jg(cm/s)

Figure 2.19 Effect on gas holdup versus gas rate of baffling between spargers
(Manqiu, 1987).

1987). In one case baffles between the spargers were provided, in the other case
there was no baffling. The arrangement with baffling gave a higher gas holdup
(smaller bubbles) than the arrangement without baffling. This suggests that
coalescence may occur when spargers are free to direct gas bubbles at each other.
Baffling between the spargers may be worth evaluating.
The typical arrangement of the sparger pipework in Western practice is
horizontal, but in Chinese practice it is vertical. The argument for a vertical
arrangement is reduced deposition area for solids. The spargers need not be located
together, Amelunxen (1985) has experimented with a second array of spargers close
to the feed port and in the Flotaire column there is the facility for a second gas
injection point.

2.3.1.3 Frother addition to the gas phase

Conventionally, frother is added to the liquid phase and diffuses from bulk
solution to the interface. Frother concentration has a stong impact upon bubble size,
as was shown clearly in Figure 2.10. Laboratory experiments have indicated that
a more efficient means of frother addition is as aerosol droplets in the gas phase
(Flint etal., 1988). In these experiments, prior to entering the sparger the gas flowed
through an ultrasonic aerosol generator containing a 0.5 to 5.0 % solution of
Dowfroth 250C. Frother consumption was measured by weighing the generator
chamber before and after the experiment, and bubble size was determined using the
technique described in Section 2.2.2.2. Comparison between liquid phase and gas

33
° Liquid phase addition
■ Aerosol addition
0.3

CLOTH SPARGER

0.4 0.8 1.2 16 2.0


Frother Consumption (mg/min/cm2)

Figure 2.20 Bubble diameter comparison between liquid phase and gas phase
(aerosol) frother addition. Conditions as in Figure 2.10 (Flint et al.,1988).

phase frother addition is shown in Figure 2.20 for a cloth sparger. Smaller bubbles
were produced by the gas phase frother addition. For example, at a frother
consumption of 1.2 mg/min/cm2 (equivalent in this case to a liquid phase frother
dosage of 20 ppm) gas phase frother addition yielded dfc=0.83 mm compared to d.
= 1.46 mm by liquid phase addition.
Addition of frother to the gas phase in an operating column is a plausible
method for yielding smaller gas bubbles. The minimum aerosol droplet size that is
required is unknown. With a high gas velocity in the gas line it may be possible to
add frother solution by aspirating or pumping. This technique may also permit the
use of spargers with larger holes, which may be less prone to plugging.

2.3.2 EXTERNAL SPARGERS

External in this context means gas and liquid (or slurry) are contacted
outside of the column and the mixture is then directed into the bottom of the column.
One method that has been widely employed commercially is that developed by the
U.S. Bureau of Mines (Foot etal., 1986; McKay etal., 1988). In this system air and
water (both at high pressure, with water often containing frother) are contacted in
a chamber containing a bed of ground quartz or ceramic spheres. The discharge
from the chamber is carried through steel pipes into the column from which it is

34
expelled through holes that are about 1 mm in diameter. The bubble size generated
by this sparger system is controlled by several factors, including air and water
pressure, water flowrate, frother dosage, and size and number of holes. Another
version of this device is now offered by Cominco Ltd.
Another commercial device is that on the Flotaire column. Here water
passes via a porous sleeve through which air is injected.
The advantages of these techniques compared with internal spargers
include: less chance of plugging with solids or precipitates; on-line sparger main
tenance; and a degree of control over bubble size (by manipulating gas and water).
The main disadvantages are the extra water entering at the bottom of the column and
the more complicated operation. If frother is required in the water then the frother
balance around a circuit of columns may need to be addressed.
Devices which have been used in laboratory/pilot columns include those
of Yoon et al. (1988) at West Virginia Polytechnical Institute and Al Taweel (1987)
at the Technical University of Nova Scotia. The West Virginia device passes slurry
(drawn from the tailings) and air through an in-line mixer back into the column.

2.4 SUMMARY OF CHAPTER

1. The gas holdup/gas rate relationship defines the flow regime in the
collection zone.

2. Columns should operate in the bubbly flow regime, where gas holdup
varies approximately linearly with gas rate.

3. Bubbly flow was modelled by adapting a hindered settling equation for


solids to estimate relative bubble to liquid velocity. It is valid for db< 2 mm.

4. Bubble diameter in a gas-water system can be accurately predicted


knowing gas holdup, gas velocity and liquid velocity; in a gas-slurry system the
additional data required are slurry density and viscosity, and bubble-particle
aggregate density.

5. Operating in bubbly flow places limits on bubble diameter, gas rate, liquid
rate, solids content, and the relative column cross-sectional area to sparger surface
area. Figure 2.21 summarises the effect of these variables on € versus J .

6. There is an optimum bubble size range 0.8 < db < 1.5 mm to maximise
superficial bubble surface rate.

7. Scaling up an internal sparger involves increasing sparger surface area in


proportion to column cross-sectional area. The type of sparger material may not
seem to be a major factor in determining bubble size, although it is important in
determining sparger life.

8. External sparger systems offer the attraction of on-line sparger mainte


nance and more control over bubble size.

35
frother cone

bubble loading, p
b

Figure 2.21 Summary of effect of operating and design variables on gas holdup versus
gas rate.

36
Chapter 3
Collection Zone: The Particle Collection Process

A mineral particle in a notation column is collected by a gas bubble


through one of two processes: 1) particle-bubble collision followed by attachment
due to the hydrophobic nature of the mineral surface; or 2) entrainment of the
particle within the boundary layer and wake of the bubble. (In a column,
precipitation of air bubbles on hydrophobic surfaces is not a significant factor.) A
key feature of the flotation column is that particle recovery by entrainment can be
virtually eliminated.
In the following, a model of particle-bubble collision and attachment is
developed. The model is shown to be in agreement with much flotation experience
and it is used to examine optimum gas rates and maximum collection rates in a
flotation column. The model is described in Section 3.2; the casual reader can skip
to Sections 3.3 to 3.5 for a description of the model implications.

3.1 AN EXPRESSION FOR THE COLLECTION RATE


CONSTANT, k.

Consider the situation of gas bubbles rising through a column of water


containing hydrophobic particles at a concentration c (number of particles per unit
volume). Collection efficiency EK, is defined as the fraction of particles swept out
by a bubble that collide with, attach to and remain attached to the bubble. Then, for
a cubic volume of water with side dimension L, the following expression describes
particle collection:

rate of particle removal = rate of panicle removal x number of bubbles


per bubble

At a gas velocity J and a slip velocity Us this expression is equivalent to:

^VJ^ i (3.1.)

Cancelling of terms gives:

iS = ("W Cp (3.1b)
dt db

37
This is equivalent to the expression for first-order rate process (provided that E^ is
independent of cj where the first-order rate constant k is given by:

1.5 J R_
kc = *-^ (3.2)

For a first-order rate process in plug flow conditions (for example the collection
zone of a small diameter long column) fractional recovery of a mineral particle in
the collection zone R is given by:

Rc = 1 - exp (- kctp) (3.3)

where tp is the residence time of the particle in the collection zone. Since R. is
particle size dependent (as shown in the following sections of this chapter) k is also
particle size dependent. For non-plug flow conditions, say d^ > 0.01, equation
(3.3) must be modified to account for short circuiting of particles; this topic is
addressed in Chapter 4.
Equations (3.2) and (3.3) provide a link between EK, k. and Rc. Therefore,
a fundamental analysis of the collection process that provides relationships between
EK and process parameters, such as particle size and bubble size, can be extended
to provide relationships between R£ (or k^ and process parameters.

3.2 A COLLECTION MODEL

3.2.1 BACKGROUND

The collection process can be studied by considering a gas bubble rising


through a dilute slurry. Collision efficiency of the bubble Ec is the fraction of all
particles swept out by the projected area of the bubble that collide with the bubble.
Subsequent to collision a particle will move along the bubble surface with either a
sliding (tumbling) motion or a bouncing motion. Particle trajectory simulations
show that particle bounce is insignificant for particles less than about 100 [J.m in
diameter (Dobby, 1984). The sliding particle maintains bubble contact until the
fluid streamlines carry it radially away from the bubble surface, unless attachment
occurs prior to this. Attachment efficiency EA is the fraction of all colliding particles
that undergo successful attachment during the time of contact.
It has long been considered that an important cause of poor coarse-particle
flotation is particle-bubble disruption by turbulence. Schulze (1977) has examined

38
the detachment process with a view towards determining an upper limit to particle
floatability. Jowett (1980) developed an estimate of critical particle size for
aggregate disruption based upon the translational velocities of the turbulent zones
in mechanical flotation machines. He arrived at critical particle sizes between 100
Hm(p =7.5g/cm3)and200nm(p = 4.2 g/cm3) for a contact angle of 60°. Below
the critical size, detachment should be minimal. Woodbum et al. (1971) based their
detachment analysis upon the tension developed in the skin of the bubble when the
bubble is subjected to a sudden acceleration in the turbulent fluid. They calculated
that the maximum strain induced in the bubble skin is proportional to dp15, and
concluded that the probability of detachment was (dp/dpmax)'-5. Their experiment
showed d maj( to be between 400 and 1000 |im for many minerals. If d is 400
Jim, then 5ie probability of detachment would be0.13 fora 100 urn particle and 0.04
for a 50 \un particle.
It is likely that detachment probability is considerably less than the above
values in a flotation column, where there is no mechanical agitation (or, alterna
tively, the critical size for disruption is higher).
Development of the collection model initially employs a single particle,
single bubble system; later the model is expanded to the real system of bubble
swarms. Particle collection is considered to occur by collision followed by sliding
over the bubble surface. It is not intended to model the complete range of particle
size; rather the focus is on fine and intermediate sized particles. Consequently,
particle detachment is not considered, and collection efficiency, EK, is given by:

E* = ECEA (3.4)

The immediate application of the model is toward understanding the collection


process occuring in a flotation column, but it is reasonable to believe that the model
offers a description of the collection process occurring in mechanical flotation
machines as well.

3.2.2 PARTICLE-BUBBLE COLLISION

The particle-bubble collision process has been studied theoretically and


experimentally by several investigators since the late 1960s: Flint and Howarth
(1971), Reay and Ratcliff (1973, 1975), Anfruns and Kitchener (1977), Weber
(1981), Schulze and Gottschalk (1979), Weber and Paddock (1983), and Jiang and
Holtham (1986). These authors based their analysis on the equation of motion of
a spherical particle relative to a spherical bubble (large compared to the particle)
rising in an infinite pool of liquid. Hydrodynamic drag will tend to sweep the
particle around the bubble, following the fluid streamlines. Particle inertia and
gravity act in acombined manner to move the particle out of the fluid streamline and
toward the top surface of the bubble.
A diagram illustrating the approach of the particle toward the bubble is

39
shown in Figure 3.1. The equations of motion in the x and y direction are

- v (3.5b)
y

where the term on the left hand side is the inertial term. Sk is the Stokes number,
given by

v*.y* an^ ux.y* are particle and liquid velocities respectively and u * is the particle
terminal velocity, all made dimensionless by dividing by the bubble rise velocity,
ub; t* is time made dimensionless by multiplying by (ub/db). Particle terminal

10

Figure 3.1 Illustration of a particle approaching a gas bubble, x, y and r coordinates


are dimensionless; at the bubble surface r* = x2 + y2 = 1.

40
velocity is calculated assuming Stokes1 equation:

In flotation the most important parameter determining the Stokes number


is particle diameter. For fine particles (e.g. less than about 30 \im) Sk is less than
0.1 and can be assumed equal to zero (the physical significance of this is that the
particle adjusts instantaneously to changes in the fluid flow). This situation is
described first; the analysis is then extended to intennediate sized particles.

3.2.2.1 Low particle inertia (Sk = 0)

The collision model of Weber and Paddock (1983) is the most comprehen
sive model to date. In their model total collision efficiency is the sum of
gravitational and interceptional collision:

Interceptional collision alone (ECg = 0) corresponds to neutrally bouyant particles


which follow the fluid streamlines exactly. Gravitational collision accounts for
particle mass. Gravitational collision alone (ECj= 0) is hypothetical; it corresponds
to particles having a finite settling velocity but zero dimension.

Ec. is given by (Weber and Paddock, 1983)

for0<Reb<300.

ECgis given by, (Reay and Ratcliff, 1973)

Er = "p ( l+_£ )2 sin26c (3.8)

where 9 is the angle (measured from the front stagnation point of the bubble) where
the fluid streamlines come closest to the bubble. For 20 < Reb <400 the following
correlation for 6c applies:

6c= 78.1-7.37 log Reb (3.9)

41
3.2.2.2 Intermediate Particle Inertia, Sk > 0.1

An analytical solution is no longer possible. Collision efficiencies for Sk


> 0.1 have been calculated by determining particle trajectories using a numerical
solution to the equation of motion (equation (3.5)) and finding the grazing trajectory
by trial-and-error.
An example of calculated collision efficiencies as a function of Stokes
number (Sk < 1) is given in Figure 3.2. Note the collision efficiencies fall
approximately midway between the trajectories determined under the assumptions
of the bubble following Stokes flow (Reb = 0) and potential flow (Re = °° ).

Collision efficiency is estimated from the following correlation:

Ec = ECo (1.63 Reb006 Sk*54 up*"°16) (Sk > 0.1) (3.10)

where ECo is E<, obtained from equation (3.6) for conditions when Sk = 0. To give
some perspective, the system d = 80 ^im, p =5 g/cm3 and d,, = 1.0 mm gives E
= 0.149 and Ec = 0.281; for dp = 40 jun E^^ 0.038 and Ec = 0.043 (Ec = 0.00155
Table 3.1 Example Calculation of Collision Efficiency

GIVEN CONDITIONS
d^ = 1.0 mm
p = 5.0 g/cm3
p,= 1.0 g/cm3
|X, =0.01 g/cm.s

CASE A CASE B CASE C

dp: 80 urn 40fim IOujti

DERIVED VALUES
Ub =11.7 cm/s (from equation (2.10) assuming e =0)

42
60

40 -

20-

0.05 0.1 0.2 0.6 1.0

Stokes Number, Sk

Figure 3.2 Trajectory calculations, E_ versus Sk and Reb. u * = 0.010 (Dobby,


1984).

when for d =10 \im). Calculation steps are given in Table 3.1.

3.2.3 PARTICLE-BUBBLE ATTACHMENT

This aspect of the collection process has received less attention than has
collision. A notable early theoretical study was that of Sutherland (1948). Attach
ment has been examined recently by: Schulze (1984), Dobby and Finch (1986b,
1987), Miller and Ye (1987), Luttrell and Yoon (1988), and Crawford and Ralston
(1988).
The underlying assumption in the analysis presented here is that attach
ment occurs when the intervening liquid (disjoining) film between the particle and
the bubble thins, ruptures and a three phase (solid-liquid-air) contact line forms. The
total time required for this process will be called the induction time t. Consequently,
particles with a sliding time t greater than tj are considered to have attached.
To calculate t the analysis applies the fluid mechanical arguments already
introduced in the collision model; no attempt is made to include surface forces
because they are not readily quantified. The induction time is user-selected in the
model; no attempt is made to predict it from first principles. For the sake of
discussion induction time is related to hydrophobicity; the larger is t., the less

43
hydrophobic is the particle. (Fora discussion of hydrophobicity and floatability see
Laskowski, 1986.)

3.2.3.1 Particle Sliding Time

The calculation of ts requires a knowledge of: a) the distribution on the


bubble surface of particle collision angles, b) the angle at which fluid streamlines
start to carry the particle radially away from the bubble, i.e. the maximum angle of
contact 6m, and c) the particle sliding velocity.

(a) Distribution of collision angles


The distribution is quantified by nfl the fraction of all colliding particles
that collide between the front stagnation point and some angle 6. This is calculated
using the trajectory model; a good approximation is

sin2 8
(3.11)

where 6 is given by equation (3.9).

(b) Maximum angle of contact, 0


8^ is calculated by determining the angle at which the radial component
of the particle settling velocity (directed toward the bubble surface) is equal to the
radial component of the liquid velocity (directed away from the bubble surface). At
0 > 6m the particle no longer contacts the bubble, unless attachment has already
occurred. A correlation between 8 , p and 8 is
m *p c

em = 9 + 81 PPP + ec (°-9009
(°-9"009 Pp
Pp) (3-12)
(c) Particle sliding velocity
Particle sliding velocity ve over the bubble surface is the sum of the
tangential component of particle settling velocity, u sin8, and the local tangential
liquid velocity. The surface of a flotation bubble is immobilized by the presence of
surfactant; therefore the fluid flow patterns known for rigid spheres are applicable.
Measurements of tangential liquid velocity as a function of distance away from the
surface of a solid sphere of Re = 290 (about the same as for flotation bubbles) and
8 = 45° are shown in Figure 3.3a (Seeley et al., 1975). (Radial distance is given in
terms of the dimensionless quantity (r*-1) where r* is radial coordinate r divided by
bubble radius; for our purposes the centre of the particle is located at (r*-1) = d /d.)
It is evident that there is a significant tangential velocity gradient on the upper
surface of the sphere. This will have a marked impact on the velocity of different
sized particles. For example, if ^ = 1.5 mm (ub = 17 cm/s) then the midpoint of a
particle with d =10 Jim will have a (r*-l) value of 10/1500 (or 0.0067) and u * of
about 0.1, while a 60 |im particle will have a (r*-l) of 0.04 and u * of about oV a

44
0.1 0.2 0.3 0.4

Radial distance, r*-1

Figure 33 (a) Liquid tangential velocity on the surface of a solid sphere versus
distance from surface (r*-l) at 6 = 45° (Seeley et al., 1975). Experimental data is for
Re = 290.

to

-0.8
>,

o
o

06

e O4

02

0.06 O2 O3 0.4

Radial distance, r*-1

(b) Model fit to data in (a)

45
four-fold increase in velocity and, consequently, a four-fold decrease in sliding
time. (This calculation has ignored the particle settling velocity component - when
including upsin0 the difference in particle sliding velocity is even greater.)
Figure 3.3a clearly shows that the assumption of either Stokes flow or
potential flow of liquid around the bubble is invalid. The velocity data has been
modelled by two linear functions (Dobby and Finch, 1986b) as illustrated in Figure
3.3b. The model makes use of the parameter surface vorticity ^ which is the liquid
tangential velocity gradient at the surface of the sphere. Some values of ^ are given
in Table 3.2 as a function of 8 and Reu.
b

Table 3.2 Some Values of Surface Vorticity

For dp/d,, < 0.03 (i.e. (r*-1) < 0.03) particle tangential sliding velocity ve,
is
d
ve = 0J^\~+ uPsine (3.13a)
<k
For dp/a^ > 0.03 (i.e. (r* -1) > 0.03) the particle velocity is calculated by
dividing the particle into two zones, the lower part that sees a velocity gradient and
the upper part that sees a constant velocity. Then vfl is given by

(d -0.03 d.) 0.03 d.


ve = 0.07 €s u,, [ —e -10.06 + ( -*) 0.03] + u sine (3.13b)
d d
p p

The particle sliding time t can now be calculated

ts = ( )TT(d +d)/(v) (3.14)


360

where 6 is in degrees and 7e is the average particle sliding velocity, determined from

46
3
0

*• 60
E_
Experimental
o values
.§ 40

I 20
(0

0 20 40 60 80
Angle of contact (degrees)

Figure 3.4 Particle sliding times. Measurements by Schulze and Gottschalk (1979);
predictions from equation (3.14) and potential flow assumption (for 6m = 90°) (Dobby
and Finch, 1986b).

equation (3.13a or b) using average values of £s and sin6.


Figure 3.4 shows a test of the sliding time calculations. Schulze and
Gottschalk (1979) measured the time that 160 u.m galena particles remained in
contact with a 3 mm diameter bubble after collision at various angles 0. Equation
(3.14) provides a reasonable fit; the potential flow assumption, used in the past, does
not.

3.2.3.2 Attachment Efficiency

A particle attaches to a bubble when the sliding time ts equals or exceeds


the induction time t. Let 6' be the angle 9 in equation (3.14) when t. = t.. After re
arrangement this gives
360 v.t.
9'= e 5_!_ (3.15)
m

Consequently, attachment efficiency is given by

sin26'
(3.16)
EA =
sin

47
20 30 40 50 60
Induction Time, t, (ms)

Figure 3.5 Attachment efficiency versus induction time and particle size, Condi-
tions: db = 1.0 mm, Ub = 10 cm/s, |i, = 0.01 g/cm-s, p = 4.0 g/cm3, e = 0

The attachment model is now complete. Figure 3.5 shows a plot of E,


versus t. for a series of particle sizes at db = 1.0 mm. At a given induction time
attachment efficiency increases with decreasing particle size. For example, if t =
20 ms then EA (80 \un) = 1.5%, and EA (<10 urn) > 99%. This strong particle size
effect is not unexpected, given the relationship between particle size and sliding ve
locity.

3.2.4 COLLECTION MODEL SUMMARY

The collection model for a single particle-bubble system, equation (3.4),


is the product of the collision model, where Ec is given by equations (3.6) to (3.10),
and the attachment model, where EA is given by equation (3.16). The implications
of this model are explored in the next section.

33 APPLYING THE COLLECTION MODEL: EFFECT OF


FLOTATION VARIABLES

3.3.1 PARTICLE SIZE AND HYDROPHOBICITY

Figure 3.6 illustrates the relationship between E and d for a given t (for
conditions see figure caption). The first feature of note is that E^ passes through a
maximum. The maximum is explained by the opposing effect of particle size upon

48
20 40 60
Particle Size, dp (em)

Figure 3.6 Collection efficiency versus particle size and induction time, from
flotation model, d = 1.0 mm, 11 = 10 cm/s, H, = 0.01 g/cm-s, p = 4.0 g/cm3, € =0
(Dobby and Finch, 1987).

collision and attachment; asd increases Ec increases but EA decreases. The peak
in EK will correspond to a pealc in collection rate constant k. (equation (3.2)). This
is illustrated in Figure 3.7 for galena flotation in a laboratory flotation column. The
rate constant peak in turn will lead to a peak (or plateau) in recovery as a function
of size, which is the common observation in most flotation systems (e.g. Trahar,
1981). The decrease in recovery with particle size above the peak has been
explained by bubble-particle detachment (Woodbum et al., 1971) or by an increase
in t with d (Jowett, 1980). The present model shows that it can be explained by the
effect of d^ on EA.
The second feature is that the peak shifts to smaller d as t increases, i.e.
as particle hydrophobicity decreases. This appears to be in agreement with
observation; for example it is commonly observed that in separating two hydropho-
bic minerals the less hydrophobic one reports over a finer size range than the more
hydrophobic one.
Finally, Figure 3.6 suggests that collection efficiency for very small
particles (d < 5-10 urn) is quite insensitive to induction time. This is because small
particles have a low velocity over the bubble and consequently a long ts; thus for
fine particles EA is insensitive to moderate changes in t. This observation has
important implications for selectivity. Consider the separation of two minerals, one
strongly hydrophobic with t. = 15 ms, the second weakly hydrophobic with t. = 40

49
ms. The relative collection efficiency with d is shown in Table 3.3. The separa
tion ranges from excellent at the coarse ancftntermediate sizes (d > 20 fim) to
virtually non-existant for dp < 5 urn. The decrease in selectivity with decrease in
particle size is well known. In mechanical cells the problem of selectivity is further
compounded by entrainment but this model suggests that at least part of the problem
has its origin in the collection process.
The assumption used in the above analysis was t is independent of d . This
is not necessary to solve the model, but no generally accepted relationship between
tj and dp in the dynamic conditions of flotation is available. The model can be used
to estimate t. by back calculation from measured recovery data. This procedure
shows t. increases as d decreases for d < 50 \un (Shaning, 1985; see also Crawford
and Ralston, 1988).

Table 3.3 Relative Collection Efficiency of Weakly to Strongly Floating


Minerals at Two Bubble Sizes

3.3.2 BUBBLE SIZE

The results in Figure 3.6 were for db = 1.0 mm. The effect of bubble
diameter is explored in Figure 3.8, where it is assumed t. is independent of bubble
diameter. Decreasing bubble size increases collection efficiency. This results from
increases in both Ec (see equations (3.6) and (3.7)) and EA. The increase in EA is
because the fractional decrease in particle sliding velocity on a smaller bubble (u
decreases with decreasing d^ exceeds the fractional decrease in sliding distance.
Experimental evidence of increasing collision efficiency with decreasing d has
been provided by Anfruns and Kitchener (1977).
The increase in EK with decreasing db is attractive. However, smaller
bubbles do not improve selectivity. Table 3.3 shows that a^=0.5 mm gives similar
selectivity to o^ = 1.0 mm (assuming still that recovery by entrainment is zero).

50
~ 0.04

5 10 20 40 80
Particle Size, dp 0»m)

Figure 3.7 Rate constant versus particle size for galena flotation in a 2.5 cm
diameter, 2 m long laboratory flotation column (Dobby, 1984). J = 0.33 cm/s, db = 1.0
mm. 8

20 40 60

Particle Size, dp ((

Figure 3.8 Bubble diameter effect on EK, from collection model.


t = 20 ms, H, = 0.01 g/cm-s, pp = 4.0 g/cm3, Ub = 2.7 cm/s at a^ = 1.3 mm; 10.0 cm/s at
, p

d^ 10 mm; 7.0
1.0 70 cm/s
/ at ^
^ = 0.7
07 mm; 4.8
48 cm/s
/ at db
d = 0.5
05 mm

51
3.3.3 PARTICLE DENSITY

Particle density has a pronounced effect on EK (Figure 3.9). With


increasing p (and t constant), F^ increases, the curve becomes progressively more
peaked and the maximum floatable size of particle decreases (especially for 0 > 5
g/cm3). Kp
The principal cause of the increase in EK is an increase in Ec due to the
increase in both gravitational collision efficiency (u * increases in equation (3.8))
and particle inertia (increasing Sk in equation (3.10)).PThe sharper peak and reduced
maximum size as particle density increases occurs because the increase in p
increases the particle settling velocity up in equation (3.13). As a result v increases'!
ts decreases, and EA decreases. Changes in ts have most impact on EA at coarse
particle sizes (Figure 3.5).
The effect of particle density on the shape of F^ vs d appears to be
confirmed in practice. Trahar (1981) reports a sharp peak at around 20-40 jtm for
cassiterite flotation (pp= 7.0 g/cm3). At the other end of the density scale, coal
flotation (pp = 1.3 g/cm3) is practised up to 800 Jim (Apian, 1976).

3.3.4 VISCOSITY

Viscosity ^ affects both bubble rise velocity and particle settling velocity,
but more importantly it likely affects induction time. It is assumed here that t. «= (0.,
following the work of Jowett (1980). Figure 3.10 has been constructed for'initial
conditions of t = 20 ms at (X, = 0.01 g/cm-s. (Also t^ «= (1, ° ^ (Jameson, 1984) was
assumed but this assumption made little impact on EK compared to assuming t. «=
I-L,). It is evident that a decrease in viscosity increases collection efficiency. This
observation may be relevant to milling operations which experience a seasonal
change in water temperature.
No attempt is made here to extend the argument to the much more difficult
case of slurry viscosity.

3.3.5 BUBBLE SWARM

To this point the model is a single bubble model. It is intuitive that there
would be interaction between bubbles in a bubble swarm. LeClair and Hamielec
(1968) have calculated surface vorticity as a function of Reynolds number and
volume fraction of spheres (or in this case gas holdup e ). From this experiment the
effect of bubble concentration on surface vorticity and, therefore, on attachment ef
ficiency can be accounted for. The assumptions are made that collision efficiency
is not affected by bubble swarm, except by the lowering of the bubble rise velocity,
and that bubble diameter is constant. Under these conditions, Figure 3.11 was
constructed. An increase in gas holdup increases EK; the principal reason is the
decrease in UL.

52
0 20 40 60

Particle Size. dp (cm)

Figure 3.9 Particle density effect on EK, from collection model, t = 20 ms, = 1.0
mm, Ub = 10 cm/s, Ji, = 0.01 g/cm-s, eg = 0

o.s

20 40 60

Particle Size, dB (pm)

Figure 3.10 Viscosity effect on EK, from collection model, p = 4.0 g/cm3, eg = 0; at
\L = 0.007 g/cm-s, t. = 14 ms, d,, = 0.91 mm, Ub = 11.9 cm/s; at (0, = 0.101 g/cm-s, t =
20 ms, ^ = 1.0 mm,' Ub = 11.8 cm/s; at Jl, = 0.015 g/cm-s, t = 30 ms.

53
0.05

20 40 60

Particle Size, dp (jim)

Figure 3.11 Gas holdup effect on EK, from collection model, d = 1.0 mm, \1 =0.01
g/cm-s, pp = 4.0 g/cm3; at €g = 5%, Ub = 10.1 cm/s; at eg = 10%, Ufc = 8.6 cm/s

Superficial gas rate, Jg (cm/s)

Figure 3.12 Example of simulated effect of gas velocity on d^, EK and k for an
operating system where db=CJg025 (symbols represent model calculation points).
Conditions: d =10 prn.p =4.0, t.=25 ms, and e from 3% to 18% (Dobby and Finch
1986c). S

54
3.4 OPTIMUM GAS RATE FOR COLLECTION

The relationship between collection rate constant and gas rate has been
given by:

kc= 1-5JgEK<V' ^
Thus, for a given flotation system operating at a specific gas velocity and bubble
size, an increase in gas velocity would be expected, from equation (3.2), to yield an
increase in collection rate. However, without altering the sparging system, an
increase in J increases db (Section 2.2.2.1) which reduces kc both directly (equation
(3.2)) and indirectly by reducing EK (Figure 3.8). The decrease in the quotient (ER/
d.) offsets the effect of increasing J . To illustrate, flotation of 10 nm particles with
a density of 4.0 g/cm3 has been simulated using the collection model. An induction
time of 25 ms was assumed. (Experimental calibration of the collection model
(Dobby, 1984; Shaning, 1985) has shown this to be a reasonable value of t. for
flotation of 10 (im particles.) As a base condition an operating system is assumed
to be generating bubbles with c^ = 1.06 mm at J = 0.50 cm/s and € = 5%; gas
velocity is then varied from 0.3 to 2.0 cm/s with resulting gas holdup varying from
3 to 18% (while assuming 0^= C J 025). The simulation result is shown in Figure
3.12. The notable feature is a peak in kc, at about J = 0.7 cm/s in this case.
There is experimental evidence which shows this peak in kc with J .
Laboratory column flotation tests that eliminated the froth zone gave a maximum
weight recovery (and therefore maximum kc) at J ~ 1.5 cm/s (Figure 3.13a). Similar
observations are made in mechanical flotation machines as shown in Figure 3.13b
where a model fit is included. Inspection of the model shows that the location of the
peak depends mainly on t.
Both of the experiments summarized in Figure 3.13 were designed to
measure effect of gas rate on particle collection only; the froth zone either was
eliminated or had a recovery approaching 100% and entrainment was eliminated as
a recovery mechanism. An analysis of the effect of gas rate on overall flotation
performance must, of course, include an assessment of froth zone performance as
a function of gas rate. Nevertheless, the fact that the collection rate constant peaks
at a certain gas rate should be appreciated in any column work.

3.5 THE MAXIMIMUM COLLECTION RATE CONSTANT

As discussed in Chapter 2 there is a maximum gas rate J max (at a given


bubble size) for the collection zone to remain in the required bubbly flow regime.
Operating at J max will, for that bubble size, maximize kc (k,^)-
Figure 3.14 shows kmax as a function of ^ for 10 jun particles and t. = 30
ms. k begins to decrease at very small bubble diameters, because J decreases
with decreasing d,,. There appears to be little advantage gained in using bubbles
smaller than about 0.5 mm. This is a similar conclusion to that reached in Section
2.2.2.7 when considering the effect of db on maximum bubble surface rate.

55
0.5 1.5 25

Superficial gas velocity, JB (cm/s)

Figure 3.13a

0.2 0.4 0.6 0.8

Superficial gas rate, Jg (cm/s)

Figure 3.13b

Figure 3.13 (a) Measured gas velocity effect upon collection zone recovery in a 3.8
cm diameter column (Dobby and Finch, 1986a). (b) Gas velocity effect upon collection
rate constant of 25 (lm galena for a mechanical flotation machine (Laplante et al.,
1983); the model uses db=l mm, J = 0.5 cm/s, t = 20 ms and n=0.53 in db=CJ ".

56
Maximizing rate constant by operating at Jgmax is, in principle, possible.
However, a collection zone at Jg max is unstable and small fluctuations in J or d^
could cause transition to churn-turbulent flow. Practical realization of kmax is
probably not feasible.
However, an operator seeking to increase kc is well advised to consider the
relatively simple option of increasing J before attempting to decrease db.
X
(0
0.2 0.4 0.6 0.8 10 12
Bubble Size, db (mm)
Figure 3.14 Simulation of maximum collection rate constant k versus bubble
diameter when J = J
g gmax
. Conditions: d =10 (ltn, p =4.0 g/cm3, t.=30 ms (Dobby and
Finch, 1986c).
57
3.6 SUMMARY OF CHAPTER

1. A model has been developed 10 predict collection rate constant k as a


function of the principal operating variables.

2. The novel feature of the model is allowance for the liquid velocity gradient
over the upper surface of a flotation sized bubble, which introduces a strong panicle
size effect into k£.

3. Rate constant shows a peak with particle diameter in accord with the
general observation.

4. Decrease in bubble diameter increases k but not selectivity.

5. An increase in particle density decreases the particle size at which k peaks.

6. For a given operating system, it is demonstrated that there is an optimum


gas rate that maximizes the collection rate constant.

58
Chapter 4
Collection Zone Mixing

The particle collection process in a column is considered to follow first-


order kinetics relative to solids concentration with a rate constant kc. The recovery
of a component from a first-order rate process is dependent upon three variables: the
rate constant, mean residence time, and a mixing parameter. Thus, the mixing
conditions within a column must be known or at least estimated for scale-up and
design. This chapter presents methods for quantifying the mixing occurring within
the collection zone of a column, and relates column dimensions to mixing condi
tions and to separation performance.

4.1 MIXING MODELS FOR FIRST-ORDER KINETICS

One extreme of mixing is plug flow transport, where the residence time of
all elements of the fluid (and all mineral particles) is the same. Plug flow in a column
will mean there is a concentration gradient of floatable mineral along the axis of the
column. The otherextreme is a perfectly mixed reactor, where there is a distribution
of retention time and where the concentration is the same throughout the reactor.
For a first-order rate reaction with a rate constant k, exhibiting plug flow transport
and having a retention time t, recovery R is given by

R = 1 -exp(-kt) (4.1)

and for a system exhibiting perfect mixing with a mean residence time T

R = 1 -(1+kT)1 (4.2)

Mixing has a detrimental effect upon recovery. For example, when t=X=5 minutes
and k = 0.5 minute'1, recovery in plug flow transport is 92% while recovery in
perfectly mixed flow is only 71%. It will be illustrated later that mixing also has a
detrimental effect upon separation.
A 5 cm diameter by 5 to 10 m high column, typical of laboratory/pilot
studies, approaches plug flow transport, while the liquid and solids in plant columns
are transported under conditions between those of plug flow and perfectly mixed
flow. Transport conditions that do not approach either of the two extremes are
usually described by one of two mixing models: tanks-in-series or plug flow
dispersion. Tanks-in-series modelling is well suited to a row of mechanical
flotation machines; for flotation columns, the plug flow dispersion model has been
shown to provide a good description of the axial mixing process in the collection
zone (Rice et al., 1974,1981). As well, column dimensions can be related directly

_ 59
to plug flow dispersion model parameters, but not to tanks-in-series parameters.
Thus, a suitable model for the collection zone of a column is a one-dimensional
(axial) plug flow dispersion model (it is assumed that radial dispersion is perfect).
Consider the downward flow of either water or mineral particles in a
flotation column. When a tracer is impulse injected at the top of the recovery zone
the concentration of tracer at a given time and at a given axial distance downstream
from the injection point is a function of the turbulent mixing with the column. The
degree of mixing is quantified by the axial dispersion coefficient E (units of length2/
time). If tracer concentration at the tailings discharge is measured with time (time
zero corresponding to the impulse injection) then a residence time distribution
(RTD) of the liquid or solids is obtained for the collection zone. RTDs can be
modelled mathematically by using two parameters to describe the mixing condi
tions: the mean residence time and the dimensionless vessel dispersion number N .
In the collection zone of height H. the vessel dispersion number for liquid or solids
is given by

where u is either the liquid interstitial velocity or the particle velocity. (In the
literature the inverse of Nd, called the Peclet number, is sometimes used.) It is shown
later that E is a direct function of column diameter, so that N can be predicted from
process and design parameters.
The objective of measuring the mixing parameters is to quantify the effect
of mixing upon recovery. The relationship is given by (Levenspiel, 1972)

2N
R=l- (4.4)
(l+a)2exp(—) -(l-a)2exp(—)

where

a = (l+4kTNd)1/2

Equation (4.4) defaults to equation (4.1) for plug flow transport (E = 0, and therefore
Nd=0) and defaults to equation (4.2) for perfect mixing (E = °o). Levenspiel (1972)
expresses equation (4.4) graphically at the same time as showing the ratio V/Vp,
where V is the volume of a real reactor and Vp is the volume of a plug flow reactor
that gives the same recovery as obtained with the real reactor. A diagram similar
to Levenspiel's is shown in Figure 4.1. The kT lines represent lines of equal volume,
or holding time, for constant k. As an example of the application of Figure 4.1 and
equation (4.4) consider the following hypothetical data. Plug flow laboratory

60
conditions (an approximation for a laboratory flotation column) are assumed to
yield a collection zone recovery (with first-order rate kinetics) of 86.5% in a reaction
time of 5 minutes. Thus, from equation (4.1) k. = 0.4 min1 and k.X = 2. If the same
reaction is to be performed in a plant column having Nd= 0.5, then employing a 5
minute mean retention time in the plant column will yield a recovery of only 75 %.
To attain the 86.5 % recovery requires a mean retention time of 8.3 minutes (VA'p
= 1.66; Figure 4.1).
Hence, the recovery is a function of k, X and Nd, and Nd is in turn a function
of E, v and H . The purpose of RTD experiments is to measure X and E; this is
achieved by fitting the experimental RTD curves to dispersion model RTDs.

4.2 DETERMINATION OF MIXING PARAMETERS

The measurement of residence time distribution has been described by


several authors (e.g. Le venspiel (1979)). A suitable technique for flotation columns
is pulse injection of a tracer at the top of the collection zone followed by sampling
and tracer analysis of the discharge. Parameter estimation from measured RTD data
requires careful attention, as discussed by Levenspiel (1979). For example, it is
essential that the tail of the distribution (i.e. at long times) is clearly identified. If
a portion of the product flows is recirculated to the feed (e.g. after another flotation
stage) then the feed must be sampled for tracer and the RTD data deconvoluted..
From the experimental data, mean residence time X and variance O2 of the
RTD can be calculated. If intermittent sampling is used (implying almost instan
taneous samples taken at times t.) then the mean and variance are given by

1 t. c. At.
T = i-l—L (4.5)
Sc.At.

and tf = S(tfT)ciAti (4.6)


2c. At.

where c. is tracer concentration at time t., and At. is given by

The relative variance O2 (= G2fX2) can be used to fit a theoretical RTD


derived from the dispersion model to the experimental RTD. When using the
dispersion model the feed and discharge boundary conditions under which the RTD
was measured must be known. Sampling from the tailings stream, i.e. sampling
outside the column, satisfies the discharge boundary condition of a closed-end
reactor. The feed boundary condition is complicated by having two feed locations

61
1.0
0.8 0.6 0.4 0.2
Fractional recovery, R
Figure 4.1 Recovery using the plug flow dispersion model, plotted as a function of
constant kt and Nd values (i.e. a graphical representation of equation (4.4)).
(the slurry and the bias water). To approximate the feed conditions of a closed-end
vessel the tracer pulse injection should be at the top of the collection zone; for
laboratory columns the top of the collection zone can be considered to be at the feed
point (see Section 4.3). With both feed and discharge conditions considered to be
closed-end, Nd is inferred by solving iteratively the following
\ = 2N. - 2NH2[l-exp(-l/Nj] (4.7)
Then, knowing u (u = H^X) and Hc, E can be calculated, using equation (4.4).
Levenspiel (1979) recommends an upper limit of N ~ 1 for model validity.
Flotation column RTDs and model fits have been reported (Dobby and
Finch (1985), Laplante et al. (1988)). Measured liquid tracer concentrations are
shown in Figure 4.2 (and Figure 5.10) for 0.45 m and 0.9 m square columns
upgrading molybdenite at Les Mines Gaspe, Quebec. The results of fitting the
dispersion model to the tailings tracer data are shown in Figure 4.3. Solid tracers
(MnO2) were also evaluated on the Gaspe columns, at particle sizes between 37 Jim
62
10 20 30

Time.tj (min)

Figure 4.2 RTD curves for liquid in the collection zones of 0.45 m and 0.9 m square
columns at Les Mines Gaspe" (Dobby and Finch, 1985). Tracer was injected just below
the pulp/froth interface.

(400 mesh) and 150 \un (100 mesh). Two of these RTDs are shown in Figure 4.4.

4.2.1 AXIAL DISPERSION COEFFICIENT

Liquid axial dispersion coefficients E, were calculated from the model fits
illustrated in Figure 4.3. Using these test results, as well as those of studies on
chemical contacting bubble columns, a linear correlation between Et and column

FDCF-F 63
0.08 OATA
MODEL -T= 13.2min
Nd= 0.476
0.06

c 0.04
o

I 0.02

o
o

a OATA
— MODEL ' 7 = 12.6 min

0.04
"A
dc. 0.45m

0.02

10 20 30
Time.tj (min)

Figure 4.3 Fitting of RTD data from Figure 4.2 to the plug flow dispersion model
(Dobby and Finch, 1985)

diameter has been observed, given by

E,= 0.063 dc(—)03 (4.8)

where dc is in metres, J is in cm/s and E, is in m2/s.


The J term in equation (4.8) is from the work of Baird and Rice (1975) on
large diameter Dubble columns. The increase in E, with increasing J results from
the increase in upward flux of slurry in the bubble wake. A complication here is
bubble diameter; over the range of bubble size relevant to column flotation (d =0.5
-1.5 mm) E( increases with decreasing db(Laplanteetal., 1988) but this has not been
quantified.
Laplante et al. have suggested E, «= dc133 and that E, decreases as slurry
density increases. However, there is insufficient data yet to confirm either of these
relationships in the case of flotation columns and it is generally not necessary to

64
have a more precise knowledge of E, than that provided by equation (4.8). As well,
Laplante et al. modelled RTD of the collection zone by introducing a distribution
of axial dispersion coefficient, and concluded that a single value of E, is adequate.
An important amendment to equation (4.8) arises from the fact that
backmixing is compounded to a very large extent by only a slight departure from
vertical alignment of the column. From a series of tracer experiments on 2 cm to
10 cm diameterbubble columns that were intentionally misaligned vertically, Tinge
and Drinkenburg (1986) have developed the following correlation

Ea = E (1 + 1100 dca)2 (4.9)

where E is the dispersion coefficient for a column of diameter dc (in metres) that
is rotated by a radians from the centre, and E is for perfect alignment. As an
example of the importance of equation (4.9), consider a 1 m diameter by 10 m high
column, and assume that equation (4.9) can be extrapolated to this case. Then a
vertical misalignment of 1 cm over the 10 m length (certainly a small misalignment)
is equivalent to a. = 0.001, and thus EJE = 4.4. Consequently, a small vertical
misalignment in the column causes a large increase in axial mixing. In clear
laboratory columns this phenomenon is revealed by a large circulatory flow pattern.
(In fact, a good guide to laboratory column alignment are the bubble and liquid flow
patterns.) The effect of alignment has not been studied in large columns.
The solids axial dispersion coefficient E is the relevant value for flotation
studies. Table 4.1 summarizes the dispersion model parameter fits for the solid
tracer tests shown in Figure 4.4. From these results it has been concluded that the
solids and I iquid dispersion coefficients in flotation columns are similar, confirming
the observations from bubble column experiments (Rice et al., 1974). Thus,

E = E= 0.063d (i)03 (4.10)


p ' c 1.6

Table 4.1
Fitted RTD Parameters for MnO2 Tracers in a 0.45 m (square) Column

65
DATA
MODEL 7. 90 mm
0.174

0 10 20 30
Time.t, (min)

Figure 4.4 RTDs of solid (MnO2) tracer for the 0.45 m column at Les Mines
Gaspe (the liquid tracer RTD in Figure 4.3 is from the same test). (Dobby and Finch
1985)

4.2.2 PARTICLE MEAN RESIDENCE TIME

The particle residence time distributions of Figure 4.4 show the expected
trend of reduced residence time with increased particle size. For a countercurrent
column operation X can be estimated by

X - (4.11)

where U. is the particle slip velocity and can be obtained from the general equation
proposed by Masliyah (1979) (a similar equation was used to describe bubble slip
velocity, equation (2.6)):

\2.7

Sp
(4.12)
18Hf(l+0.15Rea687)
where

Rep =

66
Yianatos et al. (1986a) assessed the effect on T /T, of the relative solids to liquid
upward flux due to entrainment in the wake of bubbles. The data suggest that in
flotation columns the concentration of solids in the bubble wake is the same as in
the bulk slurry, and that equation (4.11) is adequate for estimating particle retention
times. Mean particle residence times from the MnO2 tests at Les Mines Gaspe
(Table 4.1) are plotted versus particle size in Figure 4.5; the predicted values using
equation (4.11) are a good description.
To illustrate the effect that liquid velocity has upon T/X,, consider the
particular case depicted in Figure 4.6. The diagram illustrates the sensitivity of
T IX to liquid velocity, especially for particles greater than about 40 \im. This can
pose a problem in using laboratory columns that are short, say 2-5 m, where it may
be necessary to keep J low in order to attain about the same mean residence time
of an industrial 10-12 m high column.

— Predicted

* Measured

2.
20 40 60 80 100 120

Average particle size,dp(^m)

Figure 4.5 Mean particle (MnCL) residence time versus particle size, predicted
(Equation 4.11) and measured. (Doboy and Finch, 198S)

4.2.3 PARTICLE VESSEL DISPERSION NUMBER

The velocity term in equation (4.3) written for particles is u=SJ( 1 - 6 ) +


U_ . Substituting this and equation (4.10) into equation (4.3) gives the following
expression for vessel dispersion number for solids in the collection zone of colums

0.063 dc (J/1.6)0
(4.13)

67
0.8

0.2

0.4 03 1.2 1.6 2.0 22

Liquid interstitial velocity,uL (cm/s )

Figure 4.6 The ratio of particle mean residence time to liquid mean residence time
versus liquid velocity. Assumptions: p = 4.0 g/cm3; slurry is 20 wt% solids. (Dobby
and Finch, 1985) p

If the collection zone is vertically baffled into sections then d is the equivalent
diameter of a section (equivalent diameter is the diameter of the circle having the
same cross-sectional area as the section). The effect of vertical misalignment of the
column or baffles (equation (4.9) is not accounted for in equation (4.13).

4.3 DEFINING COLLECTION ZONE HEIGHT

The definition of Hc is complicated by the fact that there are two feed points
to the collection zone: the feed slurry, at some distance below the interface (typically
1 - 2 m in plant columns), and, at the interface, bias water plus solids dropped back
from the cleaning zone. The role played by the length of column from the feed point
to the interface can be assessed more clearly by considering an analysis of
concentration gradient above the feed point (Figure 4.7). The plug flow dispersion
model was used to determine the degree of mixing above the feed point for different
values of dispersion coefficient E when a column is operated with a bias of
0.3 cm/s (Laplante et al., 1988). Figure 4.7 shows that for a laboratory column,
having a very small E, there will be a steepgradient in concentration of feed material
above the feed point. This implies that the collection zone begins at the feed level,
and Hc is defined by feed level to sparger level. As a consequence, a laboratory
column has a distinct washing zone between the feed port and interface.
For a plant column, say dc =1 m, there is a very shallow concentration
gradient (Figure 4.7), reflecting the large degree of mixing. Therefore, for a plant
column it is preferable to define H£ by interface level to sparger level. The extensive

68
mixing of feed slurry above the feed point in plant columns suggests that the
performance of the column should be relatively insensitive to feed port vertical
location. It is not desirable to have the feed port too close to the interface as the
incoming slurry will disrupt the interface (especially if the feed is directed
upwards). It is not logical to have the feed enter too far below the interface, as this
risks excessive short-circuiting to the tailings. A compromise is to have the feed
point 1-2 m below the interface.

0 20 40 60 80 100
Distance above the feed level (cm)

Figure 4.7 Tracer concentration gradients above the feed level for plant (dc = I m)
and laboratory (dc = 5 cm) scale columns (Laplante et al, 1988).

4.4 EFFECT OF COLLECTION ZONE HEIGHT TO DIAMETER


RATIO

Flotation cell volume is principally determined by required retention time


and volumetric feed rate. Clearly a variety of height to diameter ratios (i.e. column
geometries) can give the same cell volume. However, a common H^ value is
about 10:1. It is of interest to determine why this ratio has been adopted. Equation
(4.13) indicates that a large H^ decreases mixing, and is the principal rationale
for vertical baffling of large diameter columns. But, varying the column geometry
causes changes not only in mixing, but also in particle velocity, volumetric bias
flowrate, and gas flowrate, all of which affect performance. The analysis in this
section is designed to illustrate these effects.
Consider a column design whereby one allows H^. to vary under the
constraint of a constant collection zone volume Vc. In this simulation assume Vc =
7.85 m3. The feed and operating conditions assumed are summarized in Table 4.2.

69
Figure 4.8 shows the effect of H^ upon bias flowrate QB, liquid interstitial
velocity u,(for QF = 600 L/min), and gas flowrate Q . As H/dcis increased both
QB and Q decrease as their superficial velocities are fixed (at" 0.3 and 1.6 cm/s
respectively). The equations that describe the terms in Figure 4.8 are given, as a
function of collection zone height, by

(4.14a)

(4.14b)

(4.14c)
Ac(l-€j (1-6)

Table 4.2 Assumed Feed and Operating Conditions For H/d Simulation
Study

Feed composition: 50% mineral A, 50% mineral B


Rate constants: k.A = 0.6 min"1, k.B = 0.02 min"1
Feed percent solids: 20 % w/w
Feed solids: p =4.0; d =65jim; Us =0.5 cm/s
Gas holdup: 15 % *
Bias flowrate: JD = 0.3 cm/s
B
Gas flowrate: J =1.6 cm/s*

* strictly, for this analysis, these values refer to the middle of the collection zone,
so they are independent of H

Equation (4.14a) shows QB decreases as H. increases. Equation (4.14c) is used in


the calculation of USp, T and Nd; for QF = 600 L/min, T and N, are plotted versus
H/dc (Figure 4.9), as is the collection zone recovery of mineral A, RcA. A
combination of factors act to decrease Nd with increasing H^d^ a decrease in dis
persion coefficient (E « dc), an increase in u(, and the increase in H . X increases
for the reasons described by Figure 4.6. These factors combine to yield an increase
in recovery (of both minerals A and B) with increasing H/d .
Grade-recovery curves are plotted in Figure 4.10 for four H^d ratios and
several feed flowrates. As a result of the decrease in axial mixing, higher H/d ratios
yield improved collection zone separation, although with progressively less impact
at H^ > 10:1. The extent of the improvement in separation will depend on the
collection rate constants of the primary mineral A and the secondary mineral B.
The advantages appear to lie in ever larger H^ ratios - QB is reduced and

70
6 8 10 12 14

Collection zone height, H, (m)

Figure 4.8 H/d simulation study: effect of column height on bias, liquid and gas
flowrates (Vc = 7.85CmJ, QF = 600 L/min) (Yianatos et al., 1988a).

H./dc
7.2 10.0 13.2 16.6
too

0.2

6 8 10 12 14

Collection zone height,He(m)

Figure 4.9 H(/dc simulation study: effect of column height on particle residence
time, particle vessel (iispersion number and collection zone recovery of mineral (Vc =
7.85 m3).

71
separation is improved. There are two ways in which Hydc can be increased and
both have important constraints. First, a single tall column can be used. An obvious
constraint in this case is the head room available. A fundamental constraint is that
as HJdc increases (with V. constant) Q decreases (Figure 4.8) and bubble loading
increases; eventually the bubbles become fully loaded and any increase in H/d
beyond this point is fruitless. A similar argument also applies if H is increased
while dc is held constant.
The second way to increase H^ is by baffling. This avoids the problem
of bubble loading, but the previous observation on the need for good vertical
alignment of the column (equation (4.9)) applies equally to baffles. The effect of
misalignment is great enough that improperly installed baffles could actually cause
an increase, rather than a decrease, in the axial mixing. This would be in addition
to the deterioration in performance if slurry and air are not evenly distributed among
the baffled sections. Baffling a column, it may be noted, will have no effect on Q ;
it is one reason against wash water addition to mechanical cells, which will require
a high QB because of the large surface area. A compromise Hydc is about 10:1 and
has led to the adoption of approximately 10 m x 1 m diameter collection zones as
standard base units (Yianatos et al., 1987b).

70 80 90 100
Recovery of A, RcA(%)

Figure 4.10 Hyd. simulation study: effect of HJ^ and QF on collection zone per
formance. For different QF, T is maintained constant for each column height by varying
V. P

72
The purpose of this analysis has been to illustrate the effects of mixing
upon collection zone performance. When column performance is controlled by the
kinetics of the collection process the mixing conditions and particle residence time
within the collection zone are a critical design consideration. The collection zone
mixing conditions become less significant, however, when froth removal is rate
limiting.

4.5 SUMMARY OF CHAPTER

1. Axial mixing of solids in the collection zone of flotation columns can be


described by the plug flow dispersion model, where the solids vessel dispersion
number is given by

0.063 dc(V1.6r
d [(V(iLJ)H
and the particle mean retention time is given by

t ■ x, ( J-/"'6') )

2. A slight departure from vertical alignment of a column, or the vertical


baffling within a column, may result in a significant increase in axial mixing.

3. A consequence of decreased axial mixing with increased height to diame


ter ratio of the collection zone is improved separation in the collection zone between
hydrophobic minerals. There are important fundamental constraints on the magni
tude of H /d which must be considered.

73
Chapter 5
Column Froths

5.1 FROTH STRUCTURE

The most obvious feature which distinguishes column froths from conven
tional froths is the addition of wash water. The wash water serves two purposes, it
provides the bias water and it provides the water necessary to overflow the collected
solids into the concentrate launder. The water flows will be quantified as superficial
rates: superficial bias rate JB, superficial water rate to the concentrate JCw and (total)
superficial wash water rate Jw. The principal interest is with the bias water.
The bias water replaces the water naturally draining from the froth. This
replacement tends to promote froth stability. One manifestation of this is increased
froth height. Adding bias water can increase froth height from less than 10 cm to
greater than 100 cm even in the absence of solids (Yianatos et al., 1986b).
The addition of wash water increases the water content of the froth, or
conversely, decreases the gas content or gas holdup. This is illustrated in Figure 5.1
which compares typical column and conventional (draining) froths.

GAS HOLDUP.

20 40 60
Overflow Level

FROTH ZONE

0.
UJ
o

I Conventional Froth
o

Interface Level

COLLECTION ZONE

Figure 5.1 Illustration of decrease in gas holdup in column froth compared with con
ventional froth; 2-phase (gas-water) system with frother (Finch et al., 1989).
Detail on the froth structure in the absence of solids has been revealed from
measurements of gas holdup, bubble size and tracer profiles. Figure 5.2 shows
bubble size increases (due to coalescence), most significantly in the region next to
the interface. Temperature profiles (the wash water was cooler than the pulp) also
reveal (Figure 5.3) that the froth is made up of at least two regions. The division
between regions appears to correspond to a gas holdup of about 74% (Figure 5.4)
i.e. approaching that expected from hexagonally close packed spheres.
The interpretation of the data is visualized in Figure 5.5. The froth zone
consists of three regions, described as: (a) an expanded bubble bed (next to the
interface), (b) a packed bubble bed above that, and (c) a conventional draining froth
above the wash water inlet.
Bubbles travel upward from the collection zone and enter the expanded
bubble bed after collision with the first layer of bubbles which define a very distinct
interface. As they enter, the bubbles have a relatively homogeneous and small size
and remain spherical. Bubble collisions against the interface generate shock
pressure waves which promote collisions up through the expanded bubble bed
region. This phenomenon seems to be the main cause of bubble coalescence in a
region where the high fractional liquid content ((1 -e ) > 0.26) makes film thinning
and rupture, the primary mechanism in conventional froths, an unlikely cause of
coalescence.

MEAN BUBBLE DIAMETER. do(mm)

t 2 3
*~~ Overllow Level

FROTH ZONE

O
tr

Interface Level

COLLECTION ZONE

Figure 5.2 Bubble size profile with froth depth (2-phase system) (Finch et al. 1989)

76
NORMALIZED TE«P€RATVn£

0.2 0.4 0.6 0-8


*• Owartlow L*val

FROTH ZONE

- lnt«rU£« LOtl

COLLECTION ZOKE

Figure 5 J Temperature profile through the froth; wash water cooler than feed
water, normalized temperature (jx plotted (equation (8.4)

T. - T,,

'F 'W

where T is temperature, i is location of interest (e.g. level in froth), W is wash water, F is


feed.

100
DATA: O li9lhM>15ixa: Jg-Uem/«; J.-OJcin/
a liothciiMptm: Jg.2)ein/«i J..0Jcm/

60 h
3J0

Bubble diameter, db(mm)

Figure 5.4 Local gas holdup versus bubble size revealing discontinuity at € ~
74%. Lines are model fits (Yianatos et al., 1986b). s
dashed line: packed bubble bed, gas holdup > 74%.
solid line: expanded bubble bed, gas holdup < 74%.

77
wash water concentrate

draining negative
bias
froth
positive
bias

packed
bubble bed
«,> 0.74

expanded
interface
bubble bed
(,<0.74 /level

bubbling
zone

Figure 5.5 Froth structure suggested from 2-phase study. (Yianatos et al., 1986b).

The packed bubble bed region extends to the wash water inlet level. The
fractional liquid content is lower than 0.26 and bubbles remain relatively spherical
but with an increasing range in size. Most bubbles move upward close to plug flow
provided the wash water is well distributed. In this region the rate of coalescence
is lower and appears to be mainly due to collisions caused by the faster rising larger
bubbles.
The conventional draining froth occurs above the wash water inlet level.
The bias here is, therefore, negative. The main purpose of this region is to convert
vertical into horizontal motion to recover the solids. Recent innovations have put
the wash water inlet above the froth (Redfeam, 1986, Amelunxen et al., 1988), in
which case no conventional froth region exists.

5.2 GAS HOLDUP: EFFECT OF VARIABLES - GAS RATE,


FROTHER DOSAGE AND BIAS WATER RATE

Local gas holdup (the gas holdup at a given level in the froth) was measured
by a combination of pressure difference and electrical conductivity readings along
the column (Yianatos et al., 1985,1986b). Local gas holdup versus froth height, or
gas holdup profiles, were determined.

78
5.2.1 EFFECT OF GAS RATE

Figure 5.6 shows typical gas holdup profiles. There is a sharp increase in
holdup across the interface. For a wide range of conditions the gas holdup on the
froth side of the interface is about 60%. From this point gas holdup increases with
froth height reaching 80% or more.
The effect of increasing J is shown schematically in Figure 5.6. As J
increases, e in the froth zone decreases. This decrease is caused by entrainment of
collection zone water (slurry) across the interface, which increases as J increases.
In contrast, e in the collection zone increases. It is evident from Figure
5.6 that a J value could be reached where e in both zones becomes the same and
the interface is lost. Figure 5.7a shows an example of loss of interface above a
certain gas rate (dependent on frother dosage); loss of interface was judged by
equivalence of froth zone and collection zone electrical conductance (see Section
8.1.1.5). That a loss of interface can occur if J is increased above a certain value
must be appreciated and is one of the reasons that an upper limit to Jg is observed
in plant practice.
A consequence of increased water (slurry) entrainment into the froth zone
is that the interface level may decrease upon increasing J . (The initial dynamic
response upon increasing J is to increase the level but the new steady state position
can be lower). When stabilizing control strategy is to control level by wash water
GAS HOLDUP £,<%>

20 40 60 80
- Overllow Level

FROTH ZONE

O
E

*■ Interface Level

COLLECTION ZONE

Figure 5.6 Schematic illustration of effect of increasing gas rate on gas holdup
profile (draining froth above wash water inlet omitted) (Finch et al., 1989).

79
3.5 1 1

.2 3.0 Frother cone (ppm)


■ IS
A 25

Loss of Interlace

J 1 J L

0 0.5 1.0 1.5 2.0 2.5 3.0

Superficial gas velocity, Jg(cm/s)

Figure 5.7 (a) Loss of interface upon increasing gas rate for given frother dosage
(conditions same as in (b)),

4)

0.0 1.0 2.0 3.0 4.0 5.0 6.0


Superficial gas velocity, Jg (cm/s)

(b) Bias rale as a function of gas rate and frother dosage, illustrating onset of negative
bias; frother was Dowfroth 250C.

80
rate the wash water rate will be increased when Jg is increased. This interdepend
ence of gas rate and wash water rate has been noted and attempts have been made
to incorporate it into an optimizing control strategy.
It is feasible that above a certain gas rate sufficient water becomes
entrained to drive the froth zone into negative bias. This will be evidenced by the
flowrate of water in the overflow exceeding the wash water rate. Just such a
condition is illustrated in Figure 5.7b; bias rate became negative above a certain Jg
(dependent on frother dosage). Stabilizing control strategies which do not measure
bias and control the level by tailings rate must respect the possibility of driving into
negative bias. This possibility imposes another constraint on the upper gas rate. The
increase in negative bias with increased gas rate above the limiting Jg is equivalent
to increased feed water recovery, which is the known situation in conventional
drained froths (Engelbrecht and Woodbum, 1975).
Increasing J will also increase the bubble size in the collection zone and
in the froth zone. In tf?e collection zone the relationship is given by equation (2.11).
The relationship in the froth zone is not known. The effect of Jg on o^ must be
considered in estimating the bubble surface area rate (see Section 2.2.2.5) at the top
of the froth zone which dictates the solids removal rate.

5.2.2 EFFECT OF FROTHER DOSAGE, OR BUBBLE SIZE

The effect of increasing frother dosage is shown schematically in Figure


5.8. The effect is similar to increasing Jg (Figure 5.6), the eg values for the two zones
approach one another. Increasing frother dosage causes smaller bubbles. The
smaller bubbles have a lower rise velocity, hence the increase in € in the collection
zone. For a given gas rate, smaller bubbles cause more surface area to cross the
interface per unit time (surface area rate) and consequently more water (slurry) is
entrained into the froth zone; hence the decrease in e in the froth zone. This
conclusion is certainly valid for a^ down to ~ 0.5mm; it is possible that with dfe «
0.5 mm (sometimes called microbubbles) the water entrained per bubble ap
proaches zero and this bubble size effect on entrainment is lost (Yoon et al., 1987).
The combination of gas rate and bubble size will determine loss of
interface and onset of negative bias (Figure 5.7). The attempt to combine small
bubbles with high gas rates to take advantage of increased surface area rates,
therefore, has limitations. It was shown in Chapter 2 that the combination will cause
breakdown of the bubbly flow regime in the collection zone. Here it is shown that
loss of interface or negative bias can occur.
Most of the experiments on which these observations have been based
used Dowfroth 250C (a polypropylene glycol methyl ether), methyl isobutyl
carbinol (MIBC) and 1,1,3 triethoxybutane (TEB). The frother was added in a range
up to 30 ppm. This gives bubble sizes in the collection zone down to about 0.6 mm
(at J ~ 1 cm/s); there is little further bubble size decrease at frother dosage above
30 ppm. It is expected that other frothers will exhibit similar general behaviour,
differing only in such details as the concentration at which loss of interface occurs,

81
for example. It is probable, as Luckie and Klimpel (1986) have shown with
conventional flotation, that frother type plays a role in size-by-size flotation rate.
Much in-plant column work involves already reagentized feeds, consequently
control over the chemistry in the column is reduced. As column circuits are
designed for new operations, the impact of frother type on metallurgy can be
explored.
gas ho

20 43

FROTH ZONE

— lftie«r*ee Level

COLLECTION 2ONE

Figure 5.8 Schematic illustration of effect of increasing frother dosage on gas


holdup profile (Finch et al. 1989).

GAS HOLOUP.C,(%)

0 20 40 60 80
.- Overflow Lovol

- interface Level

COLLECTION ZONE

Figure 5.9 Schematic illustration of effect of increasing bias water on gas holdup
profile (Finch etal. 1989).

82
5.2.3 EFFECT OF BIAS WATER RATE

Figure 5.9 shows schematically the effect of increasing JB. (The level is
maintained constant in such experiments by corresponding decreases in feed rate).
The € in the froth zone is decreased by increasing JB. However, the most important
effect \s a change in the profile; the profile becomes flatter as JB is increased and
the expanded bubble bed region (Figure 5.5) shrinks.
This flattening of the profile is the result of mixing. The plug flow rise of
bubbles in the packed bubble region breaks down and mixing and channeling
occurs, causing e to become constant. As will be shown this is detrimental to the
cleaning action of the froth zone.
Wash water should enter as a sprinkle, not as a jet, to preserve the desirable
plug flow rise of bubbles. Trying to increase JB pushes the water entrance towards
jetting. This transition to jetting must depend to some extent on the design of the
wash water delivery system for which there is no accepted standard design.
Pipework passing through the froth may cause accumulation of solids, which is
undesirable. Wash water when added above the froth can be distributed from a head
tank having a perforated base.
The depth of submergence of the wash water distributor is another
variable. In general, the deeper the distributor is submerged the greater is the wash
water split to the bias, i.e. JB/JW increases with depth. This would appear to permit
a reduced wash water rate. However, it means the overflow percent sol ids increases,
and it is known that there is a maximum percent solids for free discharge of floated
material in a column, as is also the case with mechanical cells (Lynch et al., 1981 a).
Also, reduced wash water flow to the overflow may require additional launder water
to break down the froth, resulting perhaps in an increase in total water consumption.
The effect of depth is worth exploring; Kosick et al. (1988), for example, found a
depth of 7.5-10 cm below the overflow level to be optimum.
Bringing the distributor above the froth gives the smallest JB/JW value. To
maintain target JB, therefore, Jw will have to be increased. In applications with
restricted water supplies this is undesirable. There is also some concern that with
well loaded froths the water from an overhead distributor may not penetrate, but
rather short-circuit to the overflow. The advantages of the overhead water distribu
tor are the freedom from blockage by particles and the ability to inspect on-line the
quality of the distribution.
The ideal wash water distributor should give the minimum Jw while
respecting both the minimum JB for cleaning and the minimum JCw for solids
transport over the lip. A prerequisite is uniform distribution of water over the entire
froth volume. The uniformity of the distribution should be quantified to aid in dis
tributor selection. One possibility is to use temperature as a tracer. The wash water
is often cooler than the feed pulp, consequently vertical and radial temperature
profiles through the froth can be used to monitor distributor designs. A uniform
temperature around the discharge lip is one of the indicators of uniform wash water
distribution.
Tracer studies will also reveal the degree of mixing in the froth. This

83
mixing probably increases as column diameter increases. Columns larger than 1 m
diameter are usually baffled through the collection zone to reduce mixing there; it
may be worth extending the baffling to the top of the froth zone (an example where
this has been done is on the 3.05 m diameter columns at Cuajone (Amelunxen et al
1988)).

5J CLEANING ACTION AND EFFECT OF VARIABLES

Cleaning here is defined as removal of particles which are recovered by


entrainment in the water. Most attention is focussed on fine hydrophilic particles
but it is appreciated that all particles, hydrophobic as well as hydrophilic, are subject
to entrainment.
Extensive studies on conventional mechanical flotation machines have
shown that entrained particle recovery is proportional to water recovery (Lynch et
al., 198 lb). For particles less than about 5 \im, recovery is approximately equal to
water recovery (Trahar, 1981).
In mechanical cells there is little that can be done to prevent water recovery
into the froth. Indeed it is the pulp zone which must provide the water to stabilize
the froth. Consequently a considerable effort has gone into modelling water
recovery (e.g. Bascur and Herbst, 1982). A flotation column, on the other hand, can
be operated so that the froth water is provided mostly by the wash water and feed
water recovery can approach zero.
This effect is illustrated (Figure 5.10) by impulse tracer studies performed
on full scale columns at Les Mines Gaspe. Curve 4 shows virtually no feed water
reports to the overflow. Indeed little feed water has even crossed the interface;
compare curve 3 with curve 2 corresponding to locations about 10 cm above and
below the interface respectively.
The recovery of feed water represents the boundary (worst) case of
entrained particle recovery. The more open structure of the column froth compared
to a conventional froth (Figure 5.1) probably facilitates entrained particle rejection;
the water film thickness between the bubbles in a column froth is four to five times
that in a conventional froth. Nevertheless, the study of water recovery is a useful
approach to examining entrainment.
By means of salt tracers plant and laboratory scale experiments were
conducted to assess the effect of operating variables on water recovery. The data
is described in the following sections and is presented by profiles of the concentra
tion of feed water in the froth water.

5.3.1 EFFECT OF GAS RATE AND BUBBLE SIZE

The effect of gas rate is shown schematically in Figure 5.11. As J


increases the concentration of feed water in the froth increases. At low gas rates (JS
< 1.5 cm/s) feed water concentration approaches zero at a froth height of ~ 10 cm!

84
At J > 2 cm/s feed water can penetrate 70-80 cm into the froth. Eventually
sufficient fee^l water will be entrained to produce a negative bias (see example in
Figure 5.7b) and loss of cleaning action.
Gas rate is an important variable in cleaning. If high gas rates (J > 2 cm/
s) are used then deep (>1 m) froths are necessary. At low gas rates (< 1.0 cm/s)
shallow (< 0.5 m) froths should suffice. The impact of gas rate on cleaning may
impose the practical upper limit on Jg. From pilot and plant scale experience the
maximum gas rate encountered is about 3-4 cm/s but more typically it is
1.5-2 cm/s.

10 2O 30

Tint* (rain)

Figure 5.10 Tracer concentration versus time for various locations in and around an
industrial column (Les Mines Gaspe, Yianatos et aL,1987a); Jg = 0.68 cm/s, Jw = 0.60
cm/s.
COMC. OF FEED WATER IN FBOTM WATER t%>

0 20 40 SO BO
O.ordow Le>t!

FROTH ZQHL

Intvrlac* Lavcl

COLLECTION 20KE

Figure 5.11 Schematic illustration of effect of gas rate on feed water concentration in the
froth water (froth above wash water inlet is not included).

85
Frother dosage (bubble size) is an additional factor controlling feed water
prenetration into the froth, smaller bubbles leading to increased penetration.
Frother dosage, therefore, will affect the upper gas rate for onset of negative bias
(Figure 5.7b). The above figures quoted for J refer to typical bubble sizes, in the
range c^ = 0.5 to 1.5 mm. s

5.3.2 EFFECT OF BIAS WATER RATE

The effect of JB is illustrated schematically in Figure 5.12 for conditions


of relatively high Jg (> 2 cm/s). Increasing JB does have some effect on reducing
water entrainment but it cannot completely compensate for the detrimental effect of
high J .

CONC. OF FEED WATER IN FROTH WATER (*)

0 JO JO 60 SO
.-Overflow Level

- interface Level

COLLECTION ZONE

Figure 5.12 Schematic illustration of effect of bias water rate on feed water concentra
tion in the froth water (froth above wash water inlet is not included).

Increasing JB above some value (about 0.4 cm/s in this case) is actually
detrimental. This reflects the increased mixing induced at high bias rates. Feed
water entrained as far as the well mixed upper portion of the froth will be
shortciruited to the overflow.
The split of wash water between bias and overflow depends on several
factors and is difficult to predict. The superficial flow rate of water to the overflow
JCw is given by

(5.1)

86
where e 0 is the gas holdup at the top of the froth, which can be taken as typically
80% (Fi|ure 5.6). An alternative is to estimate JCw from the solids rate and percent
solids in the overflow.
Clearly, if JCw and JB are known the wash water rate Jw can be determined
(Jw = jcw + V-
If high bias rates are undesirable what is the desirable minimum bias rate?
That some wash water is essential to realise the column's potential performance is
known (e.g. Figure 1.3), but there is evidence that JB can approach zero without de
terioration in performance (Figure 5.13). The relative lack of importance of bias,
provided JD > 0, was also noted by Clingan and McGregor (1987). The desirable
target is to operate at as low a JB as possible. By doing so this would minimize the
water entering the collection zone and maximise the retention time (capacity) of the
column. The problem is one of controlling Jfl > 0. Current bias measurements take
the difference in volumetric flow of tailings and feed (either slurry or water), which
is subject to a large relative error. The problem of bias measurement is considered
in Chapter 8.
An alternative method of quantifying the bias water is the displacement
wash ratio Dw (Egan et al., 1988), where
Dw =
D... and JD are related by
JB= (DW-D^!L = (Dw-DJCw (5.3)
Ac
Egan et al. found 1.2 < Dw < 1.5 to give satisfactory froth cleaning on fine sulfide
concentrates, with no advantage to using Dw > 1.5. This range in Dw corresponds
to0.2JCw<JB<0.5JCw.
Warm wash water has been reported to give improved cleaning, presuma
bly because the reduced water viscosity aids drainage from the froth (Kaya and
0.05 0.15 0.25
Bias rate, JB (cm/s)
Figure 5.13 Effect of bias rate on grade of concentrate (5 cm diameter column, Cu-
retreatment concentrate at Mt. Isa Mines, Espinosa-Gomez, 1987).
87
Laplante, 1988). If the water is too warm the froth could collapse with conventional
frothers.
The emphasis has been on maintaining a positive bias. It is possible that
under some conditions (e.g. a cleaner scavenger stage of a column circuit where the
concentrate is recirculated) negative bias will be adopted, the extreme being no
wash water addition at all.

5.3.3 EFFECT OF FROTH DEPTH

It should be clear from the discussion thus far that the variables interact and
that recommended values for any variable can only be general. Froth depth is no
exception.
A depth of 100 cm is typically used. This depth has the practical advantage
of damping down the enlrainment caused by high gas rates. Also deep froths reduce
the need for tight control of level. Level isoften indicated by a single pressure sensor
below the interface. This sensed level is affected by gas holdup and slurry density
changes so that significant differences between sensed and true level may exist (see
Chapter 8).
Some plant experience suggests froth depth has no significant effect on
metallurgy (e.g. Clingan and MacGregor, 1987). Given that a 50 cm change in level
means only about 5% change in collection zone volume (for a typical 10 m
collection zone height) the impact on collection zone recovery will be difficult to
detect. When high gas rates are avoided, and there is good distribution of wash
water, froth depths as low as 50 cm can be expected to maintain grade, especially
if rejection of entrained free gangue is the principal objective. In some operations
froth depth is deliberately reduced to maintain recovery when feed pulp density or
feed grade increases (Amelunxen et al., 1988; Kosick et al., 1988).
A confounding problem is that level may not be sufficiently well measured
for effects of froth depth to be revealed. If froth depth is indeed not a factor over
a wide range, an optimizing control strategy is to minimize froth depth to maximize
collection zone capacity. Exploiting this, however, will require accurate level
sensing.
The discussion to this point has emphasised cleaning, or rejection of
entrained hydrophilic particles. There is some evidence that a column froth can
show selectivity between particles differing in hydrophobicity and this may be
related to froth depth.

88
5.4 SELECTIVITY

Selectivity refers to separation between particles differing in hydrophobi-


city. Differences in hydrophobicity cause differences in the probability of particle
collection by bubbles, or in engineering terms cause differences in collection rate
constant.
Particles with the higher rate constant will be progressively selected out if
particles are subjected to repeated detachment/re-attachment events. Such events
may occur within the froth due to coalescence. Bubble coalescence results in a
decrease in specific surface area and causes particle re-arrangement which may
involve detachment/re-attachment. Coalescence does occur in column froths, to an
extent presumably dependent on operating parameters including froth depth (Yiana-
tos et al. (1986b) showed bubbles doubling in size over froth depth of ~ 60 cm) and
solids loading.
Selectivity due to detachment/re-attachment inside the froth should result
in grade profiles through the froth. There is some evidence of this. Figure 5.14
shows Mo, Cu and Fe sulphide grade profiles observed at Les Mines Gaspe
(Yianatos et al., 1988b). The column was operated with a deep froth (150 cm) and
low gas rate (1.1 cm/s) and feed water penetrated no higher than 10 cm into the froth.
The upgrading of Mo above 10 cm, therefore, appears to reveal a selectivity process
in the froth. Another test with a shallow (40 cm) froth showed no upgrading; a froth
this shallow is probably well mixed by the wash water which masks any froth
selectivity process.
In another grade profile measurement, this time on a copper sulfide
concentrate (Amelunxen et al., 1988), no grade change was observed over a 150 cm
froth depth. In this case it was likely that the froth was fully loaded with solids, so
that when detachment occurred there was no free bubble surface available for re-
attachment.
These observations concern selectivity within the froth. Some of the
detached particles will return from the froth to the collection zone, known as froth
drop back, where they will be subjected to re-collection, establishing a recycle
between the two zones. This recycle will have an impact on the overall selectivity
of column flotation.

5.5 FROTH DROP BACK

There is an increasing interest in quantifying drop back from the froth (or
its complement, froth zone recovery Rf). Measurement has proved difficult. Labo
ratory techniques include: back-calculation of Rf from measurements of recovery
with and without from (Shaning, 1985); solving for the Rf which satisfies measured
recovery from cocurrent and countercurrent column operation (Contini et al., 1988);
and, most recently, isolating the froth zone from the collection zone (Falutsu and
Dobby, 1989). Measurements in full-size plants have been made by sampling down

89
'nh water tnput

.§• 50

i Chatccpytito Pyrtu Moly

I
S 100
to

■ hterla

150 -1-4.
10 20 30 60 70 80
Grade (%)

Figure 5.14 Grade profiles for molybdenite, pyrite, chalcopyrite and silica measured
at Mines Gaspe; dc = 0.45 m (Yianatos et ah, 1988b).

the column and mass balancing around ihe froth (Yianatos et al., 1988b).
There is a notable lack of understanding of froth drop back, consequently,
only tentative observations are offered.
Drop back is size dependent, decreasing as particle size increases (Figure
5.15). A ramification is that selectivity may be related to this increased drop back
of coarse particles. Coarse particles are more likely to be locked; therefore, their
selective drop back will translate into mineral selectivity.
The overall drop back may be 50% in laboratory/pilot columns and may
be even higher in full-size columns. A tentative recommendation in laboratory/pilot
column work is to assume Rf= 50% in order to back-calculate R and estimate a
collection rate constant (see Chapter 7). The increased drop back in large columns
may be related to increased mixing in the froth. Wall effects in laboratory columns
may act to stabilize the froth and reduce drop back.
Drop back is probably dependent on froth loading. This conclusion is
partly intuitive. For example, in a lightly loaded molybdenite-carrying froth, min
eral froth zone recoveries were all greater than 50% (Yianatos et al., 1988b). At the
other extreme, froths can become so loaded that the maximum concentrate solids
rate, or carrying capacity, is reached. Under carrying capacity limited conditions
froth zone recovery becomes very small.

5.6 CARRYING CAPACITY

In Section 2.2.2.7 a carrying rate was introduced with units of mass of


solids per unit time per unit of column cross-sectional area. The analysis was
developed for the collection zone and a theoretical argument was advanced for a
maximum carrying rate. In practice it is not only impossible to realise this

90
100

100
10 20 30 40
Particle size, dp(/jn)

Figure 5.15 General dependence of froth zone recovery Rf (and drop back) versus
particle size: effect of increasing J , dc and Hf is shown.

theoretical maximum but also the measurement will take place in the presence of a
froth. The practical measure of the maximum carrying rate for a given set of
conditions is called the carrying capacity. It is a critical limitation in situations
where a large fraction of solids is to be recovered from a high pulp density feed of
fine particles. It must be allowed for at the design stage.

5.6.1 CARRYING CAPACITY MODEL

As discussed in Section 2.2.2.7 carrying rate depends on the bubble surface


area rate and the solids loading per unit area of bubble. Carrying capacity can be
expected to depend on the same variables as carrying rate, therefore an analogous
equation can be written

ltd p J
p p e (5.4)

where d^. is the bubble size at the top of the column (J is already defined at the top,
i.e. STP). The carrying capacity measured in a column is dictated by the froth since
d,^ > dfe. Carrying capacity defined this way is more suited to scale-up than the
alternative parameter sometimes used, the mass of solids per unit volume of gas C
(where Cg = Ca/Jg). g

91
5.6.2 EXPERIMENTAL CARRYING CAPACITIES

Carrying capacity can be measured for a given set of operating conditions


by determining concentrate solids rate as a function of feed solids rate. When
concentrate rate reaches a maximum this gives the carrying capacity. Feed solids
rate can be varied, at a fixed retention time, by varying feed percent solids.
The expected relationship is an increasing concentrate rate with increas
ing feed rate until the concentrate rate reaches a maximum. (This assumes that
variables such as gas rate, frother dosage and wash water rate have already been
adjusted and the system is in positive bias.) The real situation is a little more
complicated. Two examples are shown in Figure 5.16. In Figure 5.16a the feed was
fine sulfide (dg0 ~ 16 urn). The maximum concentrate rate is about 34 g/min which,
for Ac=20.3 cm2, gives Ca = 1.7 g/min/cm2. Figure 5.16b is for silica (dg0 ~ 35nm)
and shows a maximum concentrate rate of 15 g/min which, for A =5.01 cm2, gives
Ca = 3.0 g/min/cm2. °
Both examples show that at excess feed rates the concentrate rate de
creases. The decrease is a result of the concentrate becoming finer. This is
illustrated in Figures 5.17 and 5.18. Figure 5.17 shows the particle size distributions
for concentrates at three feed rates from zinc sulfide column tests. The maximum
concentrate rate occurred at a solids feed rate of 80 g/min; above this feed rate the
concentrate rate and particle size decreased. For the column circuit from which the
data in Figure 5.18 is derived, the first column is being overfed and produces a finer
concentrate and lower C than the second column.
The decrease in concentrate particle size, and consequent decrease in
concentrate rate, appears to have its origin in the preferential drop back from the
froth of coarse particles (Figure 5.15). While this is a complication, the term C is
still reserved for the maximum concentrate solids rate.
Experience with laboratory/pilot-scale columns has shown that C is
relatively independent of gas rate over for the range J > 1.5 cm/s. In that case, if
C (concentrate mass of solids / unit volume of gas) is calculated (C = C /J ) C
decreases with increasing J (Figure 5.19). This is the reason that Ca£is preferred!
Part of the lack of dependence of C onJ can be traced to the variation in
o^q with J . In the collection zone the relationship between d. and J is approxi
mately (Section 2.2.2.1) g

By observation dfc in the froth zone is also dependent on J but no relationship is


known. It can be concluded, however, that

bo g

92
100 200 300 400
Feed solids rate (g/min)
Figure 5.16 Concentrate solids rate vs feed solids rate revealing maximum concentrate
rate, or carrying capacity, Ca. Pilot scale tests for: (a) (above) Cu-retreatment concentrate
Mt. Isa Mines, dc = 5.1cm (Espinosa-Gomez et al., 1988c), (b) (below) silica, dc = 2.5 cm
(Contini et al., 1988).
20 60 100 140
Feed solids rate (g/min)
where q > 0.25. The C expression becomes modified to
(5.5)
and thus the effect of J is moderated. Figure 5.19 suggests that q in equation (5.4)
is approximately equal to 1 for Jg = 1.5 to 3 cm/s.
Figure 5.20 correlates the available Ca data (Table 5.1) against the product
ofthe concentrate 80% passing size dM, and concentrate particle density p . As the
carrying capacity model (equation (5.5)) predicts, Ca increases linearly with both
particle size and density.
When dgjj is quoted in |im and p in g/cm3 the data from Figure 5.20 gives
93
100

90

80
RJ

m 70
(A

8 60
5 50

40

30

30 50 100 200

Particle size, d p (/im)

Figure 5.17 Size distribution of zinc suifide concentrate at three solids feed rates to a
pilot column. Ca occurred at a solids feed rate of 80 g/min.

-6 *6/-10

Average particle size, dp

Figure 5.18 Selective recovery of the finest fraction in 1 st stage of a column circuit
treating Zn retreatment concentrate at Mt. Isa Mines (Espinosa-Gomez et al., 1988c)

94
0.07 _

"e
0.05 5

• 0.03

15 2.0 2.5
Superficial gas rate, Jg(cm/s)

Figure 5.19 Carrying capacity Ca versus gas rate Jg revealing lack of effect for Jg
>1.5 cm/s, while C varies inversely with J .

Table 5.1 Carrying Capacity Data from Pilot and Industrial Scale Columns
(see also Figure 5.20)

'• MIM: Mt. Isa Mines


1 U. of T.: University of Toronto

95
the following correlation

C = 0.068 d8Op
8Opp g/min/cm2
or (5.6a)
C = 0.041 t/hr/m2 (5.6b)

Equation (5.6) is an empirical model and therefore only applies within the
range of variables tested. In particular, the upper dg0 to the test data (~40^m) must
be respected. There is evidence that carrying capacity does not increase in propor
tion to dp above dp~40,i.m but rather reaches a maximum. (The existence of an upper
size limit to the carrying capacity equation is not a severe restriction since carrying
capacity only becomes a constraint for fine feeds.) C probably decreases with J
forJ < 1.5cm/s. a «
An observation from the carrying capacity data is that C is independent
of column diameter at least up to dc about 1 m. Unlike mechanical cells , the lip
length to area ratio does not seem to be a controlling parameter (until, perhaps, d
> lm). The reason is probably the good flow properties of the column froth!
produced by the wash water. As column diameter increases much beyond 1 m a
horizontal mass transport limitation may eventually constrain the carrying capac
ity. At this date, it is unknown what the uppercolumn diameter is for which equation
(5.6) is valid.
For design purposes experimental Ca values are necessary. For prelimi
nary studies equation (5.6) gives an estimate of Ca provided the feed characteris
tics (particularly dg0) and Jg fall in the range covered by the empirical model.

240

Figure 5.20 Carrying capacity versus the product of concentrate panicle size d and
density p (Espinosa-Gomez et al., 1988b). Data: 6<d8n<44 Um; 2.6< p <5 8 e/cm3-
0.025<d<1.0m. p $

96
5.6.3 SOME OPERATING IMPLICATIONS OF CARRYING CAPACITY
CONSTRAINTS

As operators seek to maximize productivity from a column, the device will


be pushed towards its carrying capacity. It is probable then that many operating
columns are at this maximum production rate. This has some important conse
quences.
A column at its carrying capacity is sensitive to mineral feed fluctuations.
For example, an increase in feed grade or feed percent solids will increase the feed
rate of floatable mineral, and since the column is at its carrying capacity this will
mean a recovery loss. An operating change such as increasing gas rate, which may
work in a mechanical cell, may not work in a column. (If carrying capacity is
increased by gas rate this probably means moving into negative bias and loss of
grade.) Increasing wash water rate and raising the level is one solution used but at
the expense of grade (Amelunxen et al.. 1988). The required control action is to
decrease feed solids rate to maintain a constant feed rate of floatable mineral, a
manoeuver not readily achieved. (As an aside, an increased feed grade requiring a
decreased solids feed rate means an increased retention time. This should not be
interpreted as reduced flotation kinetics, however, since the column is not collec
tion-kinetics controlled in this case.)
This lack of ready accomodation of feed changes through operating
variable changes throws more emphasis on the design stage. Probably the best
option is series columns, with the downstream column re-treating the tailings (i.e.
a scavenger arrangement). The circuit should be designed for the maximum
projected mineral feed rate to ensure the circuit's total carrying capacity is not
exceeded. Columns in series will mean the first is overfed and may suffer a reduced
carrying capacity as a consequence (see Figure 5.18). Also, a question with a series
circuit is what to do with downstream column concentrates, especially when
mineral feed rates drop and these concentrate grades are too low for final circuit
product. There is probably no point in recycling them to a column at its carrying
capacity. This suggests column circuits could have at least 3 stages so that the option
exists to recycle the last column concentrate to the second (underloaded) column.
These and other aspects of column circuits are further considered in
Chapter 7.

5.7 FROTH STABILITY: EFFECT OF REAGENTS AND SOLIDS

It is known that stability of conventional froths can be affected by certain


flotation reagent combinations and the presence of solids. This is also the case with
column froths.
An example of the influence of reagents on column froth stability is given
by Espinosa-Gomez et al. (1988d). Fatty acid concentrations up to 30-40 ppm

97
water

Figure 5.21 Hydrophobic particle inducing coalescence of two bubbles (based on


mechanism proposed by Dippenaar (1982)).

i"

U«ht- h*avy

Froth loading

Figure 5.22 Qualitative effect of solids loading on froth stability with particle size
and hydrophobicity as parameters.

collapsed the froth regardless of frother type. The fatty acid was added with
lannagol emulsifier in 3:1 ratio; only when additional lannagol was added could a
stable froth be formed.
Of more concern are some reports of froth instability in the presence of
solids. One manifestation is a repeating cycle of froth expansion, delivery of solids

98
and subsequent collapse. A second is formation of large, visibly underloaded
bubbles in the froth.
Solid particles influence froth stability because of their size and degree of
hydrophobicity. Consider a solid particle of diameter d which forms a contact
angle (J with the air-water interface i.e. bubble surface (Figure 5.21). The second
air-water interface when it contacts the particle will advance to establish the same
contact angle. The film thickness, in the vicinity of the particle is now dpcosp. If
d cosp is less than the critical rupture thickness this particle will induce film rupture
and froth instability.
The consequences of the analysis in Figure 5.21 are: strongly hydropho-
bic particles with p > 90° will cause film rupture regardless of their size (the two
interfaces meet at the particle equator); hydrophilic particles (P=0°C) promote film
stability. This latter observation is probably of relevance in the case of column
froths since hydrophilic particle rejection is particularly efficient.
The above analysis is for a single particle, and by extension lightly loaded
froths, and has been experimentally verified (Dippenaar, 1982). The picture is more
complex as the froth loads. The particles in well loaded bubbles provide a physical
barrier to the approach of the interfaces and to the drainage of the film water, it is
as if the film water viscosity has increased.
The effect of solids on the froth, based on the above, is illustrated
qualitatively in Figure 5.22. It is known that column froths can be stabilized to a
depth of over 1 m by wash water alone (at least in laboratory/pilot columns) in
contrast to conventional drained froths. Under heavily loaded conditions even
conventional drained froths can attain depths greater than 30 cm.
Collapse of the froth in the presence of highly hydrophobic solids may be
one reason that frother addition into the wash water is sometimes required to
stabilize the froth. It may also be the origin of the observation that a column
sometimes requires less collector than does a mechanical cell; the reduced hydro
phobicity permitting a deep column froth to be stabilized.
An important practical point is that frother dosage is better quantified as
a concentration in solution (e.g. mg/L) rather than on a g/kg solids basis. This is
essential if different feed percent solids are used as, for example, may be the case
when comparing columns with mechanical cells. Significant changes in solution
frother concentration will create changes in bubble size and leads to potentially
confusing results (e.g. Luttrell et al., 1988).
The possibility of froth collapse has been a concern in the use of columns
particularly where mineral content in the feed is low (e.g. scavenging and some
roughing applications). Recent experience at Harbour Lights Mining (Subramanian
et al., 1988) has shown columns will function as roughers with low grade feeds
(about 2% mineral in this case). In some applications low loading may mean
reduced froth depths (50 cm rather than 100 cm) have to be used.
Heavy loading of the froth may have an operational consequence. With
coarse, high density solids, it is conceivable the froth zone density may exceed the
density of the slurry immediately supporting it. Consequently, the froth, or parts of

99
it, may sink through lower density regions below. This may be a factor in the
reported observation (Kosick et al, 1988) that an increase in feed pulp density can
cause froth instability (notably breakthrough of large bubbles). Solutions have been
to decrease froth depth, increase wash water and introduce frother into the wash
water in an attempt to increase solids removal rate. These changes will increase the
water content of the froth, thus decreasing froth density and restoring froth stability.

5.8 SUMMARY OF CHAPTER

1. Part of the column wash water flows downward through the froth and
eventually exits with the tailings; this is the bias flow. The balance of the wash water
reports to the overflow. Bias superficial flow JB is typically 0.02-0.1 cm/s.

2. The froth is stabilized by the bias water, permitting deep froths (> 100 cm)
even in the absence of solids.

3. A column froth has two major regions, one near the interface where gas
holdup increases rapidly to about 74%, and an upper region where gas holdup
changes more slowly.

4. Increasing gas rates and/or decreasing bubble size may lead to negative
bias and loss of interface.

5. Gas rate is an important variable in cleaning:


a) feed water, and hence entrainment of particles, is virtually eliminated
within 10 cm of the interface if gas rates are less than about 1.5 cm/s;
b) at higher gas rates, deep froths (100 cm) are required to ensure cleaning;
c) it is possible to move to negative bias at excessive gas rate;
d) the interface can be lost at excessive gas rate.

6. Bubble size also plays a role in cleaning; as bubble size decreases the
effects of gas rate noted in 5. are intensified.

7. Bias rate has a limited effect on cleaning, provided it is greater than zero.
The following are noted, however:
a) excessive bias rates (JB > 0.3 cm/s) cause mixing in the upper froth
region and loss of cleaning.
b) bias rates should be as low as possible to minimize water consumption
and maximize collection zone retention time.

8. Deep froths, depending on the solids loading, may promote selectivity due
to detachment/re-attachment within the froth.

100
9. Recovery in the froth zone decreases with increasing particle size, i.e.
froth drop back increases with particle size.

10. The column froth exhibits a maximum concentrate solids rate, or carrying
capacity C ; the following is known:
a)a C increases linearly with the product of concentrate particle size (d^
< 40 urn) andaparticle density;
b) C appears to be independent of column diameter, at least to dc = lm;
c) when the solids feed rate exceeds that which yields Ca the concentrate
particle size distribution becomes finer and solids removal rate consequently
decreases.

11. The overflow slurry has a maximum percent solids for free discharge.

101
Chapter 6
Interaction Between Zones

The previous chapters have described in detail the characteristics and


performance of the collection and cleaning zones separately. Recovery in the col
lection zone is dependent upon collection rate constant, particle retention time and
degree of axial mixing. A practical model of the froth zone is yet to be developed
to the same extent as the collection zone model. However, certain trends about froth
zone performance are clear. The froth will attain a solids carrying limitation, which
can be predicted to a first approximation by equation (5.6). Particle drop back from
the cleaning zone to the collection zone occurs even when the froth is not saturated
with solids, and the drop back increases with increasing particle size.
Clearly then, at the interface between the two zones particle transport
occurs in both directions. The primary mode of particle transport from the collection
to the cleaning zone is through attachment to rising bubbles; entrainment is usually
of negligible importance in columns. As a result of coalescence a portion of the
particles within the froth are dislodged from the bubbles and are subsequently
transported from the cleaning zone back to the collection zone by the bias water.
Much of the drop back may occur in the first few centimetres of the cleaning zone,
where rapid deceleration of the bubble and initial bubble coalescence both occur.
The purpose of this chapter is to illustrate how this interaction between the
two zones affects overall column performance. The analysis is relevant to the
selective flotation of minerals that have different degrees of hydrophobicity, i.e.
each have a collection rate constant kc-

6.1 COLUMN ARRANGEMENT VERSUS BANK ARRANGEMENT

Consider two configurations of the cleaning and collection zones, shown


schematically in Figure 6.1. Configuration (a) represents a flotation column and
configuration (b) represents a bank of mechanical cells.
In the column arrangement, let R. be the collection zone recovery and Rf
the froth zone recovery. Thus, the overall flotation column recovery Rfc is

RR,
R. = C-L— (6.1)
RcRf+l-R

which assumes that the collection rate constant is the same in subsequent recollec
tion events as in the original collection event.
For the bank arrangement let the collection zone recovery of a single stage
be given by R . and the froth zone recovery in any stage by Rf. Then overall
recovery in a single cell R, is given by
CLEANING
ZONE

a)

F-X

COLLECTION

ZONE

Column Rcovery

RR

b)
(1-R,) (1-fl,)'

CELL 1

Bajifc recovtry. n cell

Figure 6.1 Conceptual configuration of collection and froth zones for


(a) column arrangement and (b) bank arrangement.

R.R,
Cl f
R.= (6.2)
RclRf+i-Rci

and, for a bank of n cells in series, overall bank recovery R (Harris, 1976) is

Rwn)= i-U-R,)" (6.3)


The key difference between the two arrangements lies in the retention time
in the collection zone of particles that drop back from the froth. With the column
arrangement particles rejected in the froth have 100% of the original retention time
for re-collection, while with the bank arrangement the average retention time of the

104
rejected particles in cell 2 and on is always less than the original retention time.
To illustrate the difference in performance between these two arrange
ments a simulation study will be examined. The objective is to assess the effect of
froth/collection zone arrangement upon separation of two floating minerals; there
fore, it is convenient for the purposes of comparison to consider entrainment
recovery to be zero in both arrangements.
The simulation compares a bank of 2 cells in series (each cell having
collection zone mixing conditions equivalent to a single perfect mixer) with a
column arrangement where, for consistency in the comparison, the mixing in the
collection zone is considered to be equivalent to 2 perfect mixers in series. Total
mean collection retention time T is assumed to be the same for both arrangements,
as is collection rate constant kc. froth zone recovery Rf is the same for all units. For
the column, collection zone recovery is given by the tanks-in-series expression
(Levenspiel, 1972)

r =i- (6.4)

which is substituted in equation (6.1) to obtain Rfc.

For the bank arrangement

(6.4)

which is combined with equations (6.2) and (6.3) to obtain

65 70 75 80 85

Grade of mineral A (%)

Figure 6 J2 Grade-recovery for simulation example. Rf = 50% and n = 2.


X = 4 to 20 minutes,
p

105
65

65 70 75 80 85
Grade of mineral A (%)

Figure 63 Grade-recovery for simulation example at varying R and n = 2.


Data points are simulations. '

The simulation is for a feed consisting of 50% mineral A with k =0.80


min1, and 50% mineral B with k.B = 0.08 min"1. By varying X , a range of grade-
recovery values is obtained. Froth zone recovery is assumed to & the same for both
minerals. The grade-recovery plots for Rf = 50% are given in Figure 6.2 for both
arrangements. Figure 6.3 shows the results for Rf = 30,50 and 70%, and Figure 6.4
shows the results when the mixing conditions are equivalent to 4 perfect mixers in
series and Rf = 50%.
The column arrangement provides improved separation over the bank
arrangement. In the column arrangement, separation improves with increasing
froth drop back, i.e. decreasing Rf The result is analogous to the improved
separation in a rougher-cleaner circuit when cleaner recovery is reduced (provided
the rougher has sufficient capacity to handle the increased circulating load).
It is clear from Figure 6.3 that for the bank arrangement the value of R
does not affect the separation performance (only the capacity of a given bank).
Therefore, the assumption of constant Rf in the bank arrangement is not significant
in determining the shape of the grade-recovery curve.
Note that neither entrainment recovery nor selectivity in froth zone
recovery (which may occur) have been invoked in obtaining these results. Froth
zone selectivity would improve the grade-recovery relationships for both arrange
ments.

Moving the mixing conditions closer to plug flow, e.g. by increasing


Hydc or the number of cells in series, improves the grade-recovery relationship for
both arrangements (Figure 6.4); the cause of this was illustrated in Section 4.4. Ob
viously, if the mixing is equivalent to 1 perfect mixer then the two arrangements are
the same.

106
100

95

<

1 85
o

1 80
"3

I"
§ 70
en

65

I . I i I > I . L.
60
60 65 70 75 80 85 90

Grade of mineral A (%)

Figure 6.4 Grade-recovery for simulation example with n = 2 and 4. Rr = 50%.

6.2 DEFINING RATE CONSTANTS

The collection zone rate constant kc has been defined, equation (3.1). It is
an important parameter in column scale-up and simulation. However, the measure
ment of k is difficult, because experiments must be designed which decouple the
two zones to determine the individual performance of each zone.
A common approach in flotation is to use an overall rate constant, labelled
here as kf for a flotation column. This approach treats the flotation column as a
single unit. For most situations (as will be illustrated) kfc is not a rate constant in
the true sense. However, since kfc may be the rate constant that is employed for
design and scale-up of columns, it is important to appreciate the difference between
kc and kfc.
The relationship between the two rate constants is dependent on the
prevailing mixing conditions. The analysis that follows compares k. with kfc for
the two extreme transport conditions: perfect mixing and plug flow.

6.2.1 PERFECT MIXING CASE

When the transport in the collection zone is described by a perfect mixer


with mean residence time T , the overall recovery when using kfc is given by (see
equation (4.2))

Rfc = l-(l+kfcTp)-' (6-6a)

and overall recovery when using kc is given by

107
Table 6.1 Relationship between kfc and k. for a column under plug flow
transport (t =6min)

Measured kfc Assumed Rf k k/k


' (%) (mln1) J lc

1.2

0.2

[ld+kxvjRf
Rfc = i-? - (6-6b)
[l-U+^y'jRf+i-fi-o+k;^)-1]

Equating equations (6-6a) and (6-6b) and cancelling terms, yields

kfc = kcRf (6.7)


A similar analysis has been presented by Lynch et al. (1981).

6.2.2 PLUG FLOW CASE

When the transport in the collection zone is described by plug flow with
residence time t, the overall recovery when using kf is given by (see equation
(4.1)) C
Rfc = l-exp(-kfetp) (6-8a)

and overall recovery when using kc is given by

R = [l-exp(-kctn)]Rf
k [lexpCU^+ltlCk^)] *' }

Equating equations (6-8a) and (6-8b) and cancelling terms, yields

kfc = - In [exp(kctp)Rf + 1 - Rf]


p

108
k can be considered a true rate constant (Section 3.1). However, it is clear that in
the plug flow case kfc does not have the properties of a true rate constant, since kfc
is a function of t (equation (6.9)). In the perfect mixing case (equation (6.7)) kfc
is not a function of t and has the properties of a true rate constant when Rf is
p
constant
It is evident that estimating k. from kfc requires knowledge of Rf This is
illustrated in Table 6.1 for the plug flow case. Inspection of the Table suggests at
high values of k that kfc is a reasonable approximation of k. regardless of Rf If the
condition of high k. is met the implication is that a typical laboratory/pilot column
can be used to estimate k. by measuring kfc and that this kff can be used in scale-up
work. However, the transport in the scaled-up column will be non-plug flow and
Rf then becomes important (equation (6.7)): it cannot be assumed that kfc from a
plug flow pilot column is the same as kfc in the scaled-up column even when k. is
the same in both.
At present there is limited data on the magnitude of Rf or the factors which
influence Rf. It is believed that Rf in laboratory/pilot columns is typically in the
neighbourhood of 40 to 80%, and that it can be considerably lower than this in larger
diameter columns.

63 COCURRENT COLUMN OPERATION

Flotation columns have been operated with cocurrent gas-slurry or gas-


water flow (Furey, 1986; Soto and Barbery, 1987; Contini et al., 1988). Cocurrent
operation has attractions in flotation of: coarse particles (e.g. > 100 nm) where
particle settling rates give low retention times in counter-current columns (see
Figure 4.5); and, minerals with high rate constants (e.g. coal) where downward
slurry rates are restricted by gas holdup constraints (see Figure 2.12).
Interaction between the zones may now have quite a negative impact.
Particles dropped back from the froth no longer have the original retention time for
re-collection, indeed the cocurrent design may take the dropped back particles
directly to tailings (Contini et al., 1988). This means reduced recovery and loss of
the re-collection events that promote selectivity. Cocurrent columns may have to
be run with shallow froths to achieve recovery.

109
6.4 SUMMARY OF CHAPTER

1- The physical arrangement of the collection zone and the cleaning zone has
an effect on the interaction between the two zones. Separation between two floating
species (i.e. selectivity) is shown to be greater when using a column arrangement
compared to a bank arrangement.

2. The difference in selectivity between the two arrangements is because


particles rejected from a column froth have 100% of the original retention time for
re-collection.

3. Relationships between collection zone rate constant k and overall rate


constant kfc are developed for the perfectly mixed and the plug flow cases. For the
plug flow case kfcdoes not have the properties of a true rate constant.

4. In a laboratory/pilot column operating under near plug flow transport


conditions, kfc is not the same as kfc in a scaled-up column (where mixing is greater)
even when k is the same in both.

110
Chapter 7
Column Scale-up

In this chapter the scale-up methodology is reviewed, including a step-


wise calculation and a description of the test procedures for parameter estimation.
Two examples of scale-up are taken from amenability testing and parameter
estimation through to plant results. The chapter finishes with a brief review of
circuits.
There have been two approaches to scale-up using kinetic data: (a) using
an overall rate constant kfc, or (b) using a collection zone rate constant kc along with
a value for froth zone recovery Rf (The relationship between kfc and kc has been
described in Section 6.2.) The later is the more comprehensive. Both approaches
are described in this chapter.

7.1 SCALE-UP MODEL

Recovery models of the collection and cleaning zones and their interac
tion have been presented in detail in previous chapters. Their use in scale-up will
be considered in this section.

7.1.1 PARTICLE COLLECTION

Collection zone recovery is given by equation 4.4

4aexp(2N>
R = 1 - , - ^ (4-4)
1) °)
where ,„
a = (l+4kcT.Nd)1/2

= 0.063 dc(V1.6)
[(J/(l) UlH

(4,2)

Sp 18nf(l+0.15Rep0687)

Re =
e dPUsPP.(1^ (4.12a)
Mr
111
-T. (4.11)

Hd-6)
and

There is as yet no detailed model of froth zone recovery Rr but in the more
comprehensive simulation it must be considered. Overall recovery Rf is then given
by equation (6.1) c

RRf
Rfc = (6.1)
RRf+l-R

Table 7.1 Data to Illustrate Scale-Up of Collection Zone

Characteristics of Feed
Solids rate, M FS
= 25t/hr
Solids density = 4.4 g/cm3
Percent solids = 35% (psl = 1.37,4V
0.109)
.-.QF = 869 L/min
= 774 L/min
= 95 L/min

Mineral content = 50% mineral A,


50% mineral B
Mineral rate constant, k = 0.3 min'
= 0.012 mm1
Mineral particle size, dm = 37 urn

Characteristics of Column
Base unit diameter, d cited = lm
= 0.785 m2
height, Hc = 10m
Bias rate, JB = 0.1 cm/s
Number of base units = N
Gas holdup, e = 15%
Gas rate, J = 1.6 cm/s

Target Performance
Recovery of mineral A >= 70%
Grade > 88% mineral A

112
The use of this model for scale-up will be illustrated by a hypothetical
example Table 7.1 gives the necessary information. The approach adopted to
select the size of column in this example is to consider a base unit column of d
= lm and determine the number, N, of base units in parallel required to attain the
target recovery (70% of mineral A in this case). This example will use approach
(b)(useofkcandRf).

Following is a step-wise solution.

Step 1. Select N. This is the user's choice; let N = 4.

Step 2. Calculate Js|. To start Js| is not known so an estimate is required to initiate
the solution; let

where R^ is the total mass recovery, which is estimated; if the calculated


R deviates significantly from the estimate an iteration is performed.
Using an estimated R^, of 45%,

Js| = 0.54 cm/s.

Step 3. Calculate t,

T, = [10(0.85)/0.54]x (100/60)

= 26.2 min
Note the 100/60 factor is to convert m to cm and seconds to minutes.

Step 4. Calculate Us (trial-and-error using equation 4.12)


For d = 37 |im, Us = 0.16 cm/s (assuming |is, = 0.01 poise).

x 100

= 0.79

Note that the 100 factor is to convert cm to m.

113
Step 7. Calculate a, in equation(4.4), and then calculate R and R
For mineral A cA cB
a = 4.56 and R =93.8%
For mineral B
a= l.34andRD = 20.7%

Step 8. Calculate Rff using equation (6.1). To do so requires a value for R The
evidence points to Rf in industrial columns of less than 50%. Table 7 ■>
summarizes column performance for a range of Rf values.

Table 7.2 Predicted Performance for the Example as a Function of R


(RcA = 93.8% RcB = 20.7%) f

3 RfcA RfcB Rrol Concentrate Grade


(%> (%> (%) (%) (% mineral A)
20 75.2 5.0 40.1 93.8
35 84.1 8.4 46.3 90.9
50 88-3 11.2 49.8 88^5

The target recovery (RfcA = 70%) therefore is attained at an acceptable


grade within the expected range of Rf

Step 9. Iterate with the current estimate of JsJ. In this case RTo| is approximately
equal to the first estimate and consequently iteration is not necessary.

Step 10. Change N until the target recovery is reached (unnecessary in this
example).

The selection of 4 base units of dc = lm gives about the required capacity.


However, the recommended design would be a large column baffled into 4, each
baffled section being 1 m equivalent diameter. The design column diameter is given
by

c (design) ^N dc(base)
In this case, therefore

= 2m.

(It is appreciated that a circular cross-section column does not give baffled sections
which are circular, within the accuracy provided by the model this is not considered

114
important. If square columns were selected the baffled sections would retain
geometric similarity.) ,
On inspection several factors emerge as having the most impact on the
scale-up- the flowrate, x/x. and rate constants. In many situations the flowrate, Q
is not accurately known" or is known to vary widely due to changes in both solids
rate and percent solids. Such uncertainties in QF can dwarf such considerations as
the true value of the rate constants.
Thex/x factor is to allow for the settling of particles (this is a significant
distinction from mechanical cells where such allowance is not necessary). In this
case x At = 0 80 for dg0=37 ^m. If, in this example x, were used rather than xp, then
a smaller column would seem sufficient to reach the target recovery. By selecting
d to calculate x emphasis is placed on the coarser particles in the distribution.
*° For this example collection zone rate constants for minerals A and B were
used. If overall rate constants are used, then the calculation routine includes one less
step (Step 8) and column recovery Rfc is then given by Rc in Step 7. Rate constant
measurement is addressed in Section 7.2.3.1.
The choice of dc = 1 m is based on the fact that such size units have
a proven record of succes's. Other diameters can be selected. In the present case it
would be worth evaluating N = 1, dc=2m to determine if baffling has much impact;
grade as well as recovery should be considered.

7.1.2 CARRYING CAPACITY

As with rate constants, this value is best determined experimentally.


However, for calculation purposes assume it is given by the equation

Ca = 0.041 dg0 pp t/hr/m2 (5.6b)

with dg0 in Jim and p in g/cm3. Therefore, in this example


C =6.7tyhr/m2.
a

Following is a stepwise solution to estimate dc-

Step 1. Estimate concentrate solids rate, MCs.

MCs =MfsxRt
= 25x0.43

= 10.8tonne/hr

Step 2. Estimate Ac.

= 1.60 m2

115
Step 3. Estimate d.. Assuming a circular column d. = 1.43 m.

Since the d. estimated from carrying capacity considerations (1 43 m) is


less than that estimated from mixing considerations (2 m) the selected column is d
= 2 m. c

7.1.3 SPARGER SURFACE AREA

If porous spargers are to be used in the full size column the required
sparger area is estimated using the recommendations from Section 2.3.1 1 In this
case, assuming the pilot column of 5.08 cm2 (A = 20.3 cm2) used a stainless steel
sparger of surface area (As) 60 cm2, i.e. Rs ~ 3, then the recommended sparger
surface area in the full size column is 3 x (irdc2/4) = 9.4 m2 (obviously this does not
apply if external spargers are used).

7.1.4 CONCLUSIONS FROM SCALE-UP EXAMPLE

The recommended column diameter is 2m, baffled into four sections with
a collection zone height of 10 m. Allowing for froth depth (I -2 m) and space for the
spargers (1/2 - 1 m) the full height will be about 13 m. Sparger surface area is about
9.4 m2.
In this example the target performance was achieved in a single baffled
column. This is not necessarily recommended since circuits, particularly with
recycle, have inherent advantages in terms of separation efficiency. If d is
calculated to be larger than about 2.5 m (the current maximum diameter in useVor
an unbaffled column) more than one column would be recommended. By combin
ing into a network of columns circuits can be simulated, where each column can be
individually sized (Yianatos et al., 1987b and del Villar et al.,1988).

7.2 LABORATORY/PILOT COLUMN TESTING

The parameters necessary for the scale-up model (primarily the rate
constants and carrying capacity) can be obtained from laboratory/pilot-scale
columns (e.g. dc = 2.5 to 5 cm, H.= 2 to 10 m). However, before this it is usually
necessary to establish that column flotation offers improved metallurgy compared
with mechanical machines (factors such as capital and operating costs and ease of
control notwithstanding). This has been called amenability testing and will be
considered in this section also.

7.2.1 THE PILOT COLUMN

A typical set-up is shown in Figure 7.1. Typical dimensions are listed in


Table 7.3. The choice of dc = 5.1 cm (2 in) is largely from the recommendations of

116
Wheeler(1983 1985). Others have suggested that the minimum dc should be 10 cm
or even larger For testwork on limited samples, for instance testing that requires
batch laboratory preparation of the feed, smaller columns will be necessary;
columns of d =2.5 cm, H = 2 m have been used successfully for amenability and
parameter estimation work. (As a guide, the the minimum feed slurry volume
should be about 3 to 4 column volumes to ensure reaching steady state.)
Table 13 Typical Pilot-Column Set-Up (see also Figure 7.1)

diameter 5.1 cm
.\Ac= 20.3 cm2
total height* 4 to 9 m

feed position 2 m from top


sparger
stainless steel, As = 60 cm2
pumps
peristaltic, capacities ~ 0.2 to
2 L/min; controls are best
located at the level near the
feed
feedtank mixer, sample port
flowmeters gas, wash water and launder
water, these should all be
located near the feed for con
venience
wash water distributor perforated copper pipe

* e.g. 3 to 6 sections each 1.5 m for portability

Water

Figure 7.1 Typical pilot column set-up.

117
Table 7.4 gives an example of column settings for a nominal liquid
reS.dence tune xlR<m of 15 min (T|nom is the residence time if e is ignored). The feed
pump may be set at the approximate rate by discharging to atmosphere- the setting
is approbate because pumping into the column is against a head generated above
the feed port. The tailings pump may be set approximately by first filling the column
with water; it should be reset after the column fills with slurry and be checked
periodically during the run. (Note the feed rate is more difficult to check while
running.)

Table 7.4 Example of Settings for t, nori = 15 min and dc = 5.1 cm

Step 1. Volume of collection zone. Assume up to the feed port


(see Section 4.3).

Step 2.

Step 3. SetQF. Assume JT F = 0.2 cm/s


•Jp= V-Jt-f
= 0.74 cm/s
.\QH= 1.9 L/min.
(The true bias JB can only be calculated at the end of the
test from a water balance.)

Step 4. Set the wash water rate, Qw.


From an estimate of the concentrate production rate, an
estimate of concentrate water rate, QCw, can be made.
Therefore, for JB = 0.1 cm/s (QB = 0.12 L/min), initial
setting for Qw is
Qw = 0.12+ QCw L/min

An initial wash water rate can be set prior to the test by estimating the bias
QB (say 0.2 cm/s) and adding the expected flow of water to the concentrate (X. .
While a test is in progress it is important to ensure that the column is operating close
to the desired bias. Bias flow can be estimated during the run by taking a timed
sample of the concentrate (with allowance for launder water), estimating the water
fraction of the concentrate sample and subtracting this from the wash water addition
rate (measured with a rotameter).

118
Operation requires a steady froth depth, usually between 0.3 and 1 m. This
may be controlled by manipulating tailings rate or feed rate (both giving a fast
response) or wash water rate (relatively slow response). With good quality
peristaltic pumps, pilot column operation is characteristically steady and manual
control is quite feasible. Occasionally the interface is difficult to observe, especially
at high gas rates or with feeds containing a large amount of floatable solids. To
overcome this frustration will mean sensing and controlling level, a problem dis
cussed in Chapter 8. For extended pilot runs, especially involving more than one
column, automatic control is almost essential.
A period of about two residence times (e.g. 30 min for the example in Table
7.4) should be allowed after a change in conditions prior to sampling. Timed
samples should be taken of the two products and wash water and launder water rates
recorded
r
(for water balancing). A grab sample of feed (through its own sample port
next to the feed port, to minimize disturbance in feed to the column) can also be
obtained. Percent solids and chemical analysis of the feed and two products
completes the water and solids balancing. Size-by-size analysis is often warranted.
It may be possible to pump from the process stream being tested, either
directly into the column or via the feed tank (tank overflow being returned).
Commonly, however, batch samples are retained in the feed tank. In some cases
mixing in the feed tank may induce changes in the mineral surface (e.g. oxidation
of sulphides). Frequent replenishment of samples in a small feed tank may be
necessary to minimise the problem. For the same reasons, only under exceptional
circumstances should slurry samples be sent out for testing.

7.2.2 AMENABILITY TESTING

This involves establishing grade-recovery curves (or their equivalent e.g.


ash-yield in coal preparation) and comparing them against existing standard
laboratory test results, or better, against plant performance. The curves should be
established for the expected range of feed conditions (feed grades, percent solids).
The experimental variable of choice in the column is to alter either retention time
or gas rate (or both), with reduction in percent solids as an additional option if the
system is carrying capacity constrained. In some cases collecting and reprocessing
the tailings may be necessary to reach the required retention times.

A few examples will best serve at this point.

Column vs laboratory mechanical cell: Incremental time batch flotation tests in a


laboratory mechanical cell are used to generate a grade/recovery curve for compari
son with that produced in a column. Laboratory mechanical cells often produce a
superior grade/recovery to that of full size mechanical cells when compared in an
open circuit basis. Thus, if the column gives superior results to those from the
laboratory mechanical cell on a particular stream there is an implication the column
is better suited to that stream than conventional cells.

119
10 • cetunw (at ?5\ toWa k\ U
A Uboritory (at 1»% »ofid» in tt«d> l»*4 94% Cu ^

40 50 60 70 80 90 100
Recovery of Cu (%)

Figure 7.2 Cleaning of Cu retreatment concentrate (Mt. Isa Mines).

40 50 60 70 80 90 100

Pentlandite recovery (%)

Figure 7 J Cleaning of nickel scavenger concentrate (Falconbridge).

Two examples are given in Figures 7.2 and 7.3; in Figure 7.2 the column
outperforms the laboratory mechanical cell, while in Figure 7.3 the column
struggles to match the mechanical cell result. The result in the latter may indicate
that entrainment is less of a factor compared with particle locking or difficult
selective flotation. In some sulphide separations it is possible that the pulp potential
and oxygen content of the pulp in the column differs sufficiently from the laboratory
(and maybe even from top to bottom in a pilot column) to affect results. Whatever
the cause, Figure 7.3 emphasises that amenability testing is an essential first step.
Column vs existing plant: The next two examples will serve as case study
illustrations of the scale-up procedure from amenability testing through to plant
installation. The first example is the bulk Pb/Zn middlings flotation at Mt. Isa
Mines' lead-zinc concentrator (Figure 7.4 and Table 7.5a). The second example is
copper-nickel separation at Falconbridge's Strathcona concentrator (Figure 7.5 and
Table 7.5b).

120
Bulk Cone.

(a)

273SI»go
counter-curront
I g»«nini I

Rogrlnd

Bulk Rough err Tails

Bulk Cone.
(b)

Tails

Figure 7.4 Bulk Pb/Zn circuit at Mt. Isa Mines' Pb/Zn concentrator (locally bulk
Pb/Zn is referred to as low grade middlings, LGM): a) original circuit (1986) and b) cur
rent circuit (1988).

Cu Rcaflftw I lo

CuBwigtn

1.... .:

CuCtnc

Figure 7.5 Cu cleaning circuit at Falconbridge's Strathcona concentrator: a) current


circuit, b) column test circuit and (dashed lines) full scale circuit currently being
commissioned.

121
Table 75 Data on Streams For Case Studies

(a) Mt. Isa Mines Pb/Zn Middlings

Solids Rate (t/h)


Percent Solids
Feed: %Zn
%Pb
ZnS Rate Constant,
kfc (min1)

Ca (g/min/cm2)
Target Metallurgy
Grade (%Zn + %Pb)
Recovery (%Zn)

* The feed to the columns is not the same as originally tested because of the new
circuit arrangement (Figure 7.4)

(b) Falconbridge Cu-Ni Separation

(Conditions apply to 0.91 m diameter pilot column)

The amenability test results are given in Figures 7.6 and 7.7 respectively.
In both cases the column offered improved separation (in the case of Mt. Isa,
particularly on Zn; Pb metallurgy was similar in both the column and the existing
plant, although the flotation rate of galena was reduced compared to sphalerite in
the column.)

122
20 30
20 40 60 80 100
Recovery of Zn (%)

Figure 7.6 Amenability test results (Zn grade vs recovery) for bulk Pb/Zn cleaning
(Mt. Isa Mines). (Pb metallurgy in the column was the same as that in the existing plant;
grade was a consistent 10-13% at recoveries up to >90%.)

z 3 -

I 2
o
o

40 50 60 70 80 90 100
Recovery of Cu (%)

Figure 7.7 Amenability test results (recovery of Ni vs recovery of Cu) forCu-Ni


separation (Falconbridge).

7.2.3 PARAMETER ESTIMATION

7.2.3.1 Rate Constant (see also Section 7.3)

While the desirable rate constant is that of collection kc, usually the overall
rate constant kf (i.e. that obtained in the presence of a froth) is measured. The
remarks here assume that the variables gas rate, frother dosage and collector dosage,
all of which affect the rate constant, have been previously set. In the case of gas rate

123
it is convenient to work near the middle of a range where metallurgy is not strongly
affected. For example, from Figure 7.8 the recommended J would be in the range
1.5 to2.5cm/s. 8

100

a? 80 ■

0 12 3 4
Superficial gas rate, J^cm/s)

Figure 7.8 Copper recovery versus gas rate on Mt. Isa Cu retreatmeni concentrate.

Measurement of k.
— fe

Two methods of estimating kfc have been used: (a) varying tailings rate in a long
column, and (b) collection and recycle of tailings using a short column.
Method (a) - Varying QT in a long column (e.g. H. = 10 m): Typical long column
dimensions are those in Table 7.3 This column has a collection zone length of about
9 m. The maximum retention time depends on the minimum tailings rate achiev
able, which in practice is about JT=0.4 cm/s (or QT ~ 0.5 L/min for a 5 cm diameter
column). This gives a nominal xmax of 35 minutes. Longer times require collecting
and reprocessing the tailings which is similar to method (b).
In a long column the bubbles may become loaded at high feed rates, which
is undesirable (it is assumed in the first-order flotation model that the bubbles are
not loaded). Feed dilution (using the same process water) is then required to avoid
the problem.
Method (b) - Recycle to a short column (e.g. Hc = 2 to 4 m): Reasons for using a
short column include lack of head room and limited feed volume. In this method
the tailings have to be collected and reprocessed as many times as necessary to reach
the target retention time. Compared with method (a) bubbles have less time to fully
load. Disadvantages are the more frequent handling of the tailings and the fact that
transport m a short column deviates more from plug flow.
For both methods kfc is estimated from the slope of In (100-R) vs t where
R is percentage cumulative recovery and t is time. Figure 7.9 is an example of
method (a), from the Mt. Isa bulk Pb/Zn middlings case. In the 2-column test there

124
is a downturn in the curve caused by loading of the bubbles with fine particles in the

first column. Diluting the feed and processing in a single column gave slopes similar
to that in the 2nd column in the undiluted case. Over a series of tests, the range in
k. was estimated (Table 7.5a).
Figure 7.10 is an example of method (b), from the Falconbndge Cu-Ni
separation example. Particle settling rate and gas holdup (assumed to be 15%) were
accounted for in calculating retention time. The measured k(c values for chalcopy-
rite, pentlandite and gangue (including pyrrhotite) are given in Table 7.5b. In some
cases, kfc is estimated from a single point corresponding to the target recovery (see
Figure 1.9).

100

0 10 20 30 40 50 60 70
Time (min)

Figure 7.9 Rate plot for bulk Pb/Zn example (using nominal retention time).

5 10
Time (min)

Figure 7.10 Rate plot for Cu-Ni separation example (using panicle retention time).

125
Measurement ^

Collection zone rate constant kc can be calculated in two ways: (a) from k , and (b)
eliminating the froth zone. fc
Method (a) - From kfc: This approach uses the equations developed in
section 6.2. If the transport in the pilot column can be assumed to be plug flow then
an estimate of kc is obtained from equation (6.7) by using the measured k and a
value for R^ Rf = 50% is a reasonable choice for a laboratory/pilot column (Figure
5.13). If the transport in the pilot column cannot be assumed to be plug flow, then
an estimate of kc is obtained by taking the average of results from equations (6.7)
and (6.9). In the Falconbridge Cu-Ni separation example k was estimated from k
assuming plug flow (i.e. using equation (6.7)) and R. = 50%. The results are given
in Table 7.5b.
Method (b) - Eliminating the froth zone: If the froth zone can be eliminated
while still maintaining a low level of recovery by entrainment, then the measured
values of kfc will be equal to k. This can be achieved by operating at a very high
bias (> 0.4 cm/s) which effectively eliminates the froth zone, by having the water
addition well above the feed port, and keeping J <1.5 cm/s.

7.2.3.2 Carrying capacity

The carrying capacity is usually estimated by altering the feed percent


solids (to alter feed solids rate) at a given retention time until the concentrate solids
rate reaches a maximum. As in the case w ith measuring rate constants, it is assumed
that the desired operating conditions have been set; with regard to C, the imponant
two variables are gas rate and frother dosage.
Two examples of Ca measurement, one for copper retreatment concentrate
from Mt. Isa Mines and the other for fine silica, were described previously (Figure
5.16).
In the Mt. Isa Pb/Zn example, concentrate production rate was relatively
independent of feed rate, a typical finding for fine particles. The measured C was
about 4.5 g/min/cm2 (Table 7.5a) in agreement with equation (5.6).
In the Falconbridge Cu-Ni separation example concentrate rate increased
with feed solids rate. The measured Ca was about 7.5 g/min/cm2 (Table 7.5b). This
C, is less than that predicted using equation (5.6) which may be because of a large
froth drop back with such coarse particles (see for example Figure 5.15). This
emphasizes the need to measure Ca if at all possible.

7.3 LARGE PILOT COLUMN TESTING

Pilot columns of dc = 0.2 to 1 m are frequently constructed as an


intermediate stage in column design work. These large pilot columns serve to check
the amenability result and the preliminary scale-up. In addition they provide

126
Table 7.6 Typical Operating Conditions of Columns on Bulk Pb/Zn Middlings,
Mt. Isa Mines (May 1988)

COLUMN JT JT-F Hf
(cm/s) (cm/s) (cm/s) (min) (cm) (g/min/cnr)
(cm/s)

0.35/0.65 0.26/0.31 O.O2/O.O5 0.58/0.96 30/52 100/200 2.6/3.2


1
0.44/0.75 0.20/0.26 0.10/0.18 0.60/0.96 26/45 100/120 1.4/2.7
2
0.57/1.04 0.16/0.26 0.06/0.29 0.64/0.96 20/35 80/100 0.8/1.5
3

* Cr is concentrate solids rate

operator training and give an opportunity to evaluate practical aspects such as gas
sparging and process control instrumentation.
Rate constants can be extracted from the test results. The approach is to
use the scale-up model and to find by trial-and-error kfc or k. which give a fit to the
measured performance data. In the case of k^., Rf must be set first. Rate constants
derived in this manner are inherently more reliable for scale-up because of the larger
scale of testwork compared with laboratory/pilot columns.

7.4 SCALE-UP CASE STUDIES

7.4.1 CASE STUDY 1 - BULK Pb/Zn MIDDLINGS CLEANING (MT. ISA


MINES)

The selected design was an open scavenger circuit of 3 unbaffled columns,


2.5 m diameter by 13 m high. Some of the simulation work on which the selection
was based is shown in Figure 7.11. Two points of interest in Figure 7.11 are: the
target metallurgy is just met in two columns but comfortably met in three; baffling
was judged not to offer sufficient gain to warrent the additional cost and the problem
of feed and gas distribution. Not evident in the Figure is that the first column was
consistently at its carrying capacity.
The simulations were run using kfc (partly because at the time (1986) little
was known about Rf). It was felt that since the system was partially carrying
capacity limited the error in incorrect modelling of the collection kinetics would be
less of afactor. The carrying capacity limitation probably contributes to the reduced
impact of baffling which only influences collection zone performance.
The circuit was commissioned in late 1987 and the required capacity was
comfortably met; typical performance is shown in Figure 7.12. The new circuit has
contributed about a 2% increase in Zn recovery to the bulk concentrate with similar
Pb recovery as before.
The new circuit has a different configuration to that originally tested
(Figure 7.4), hence the feed characteristics to the column are different. This

127
45
55
Celurcn 1

40

8 baf Had (6 nctljni)


unbadlod 1.2*3 ~
N 35
"o

f 30 40 ■§
CO
O

25
35

20 30
0 20 40 60 80 100
Recovery of Zn (%)

Figure 7.11 Simulated effect on Zn metallurgy of number of column stages in series


and baffling of each column. All columns are 2.5 x 13 m.

45 55

40 50 £
a

35
45 ^

30 40 •§
o

25
35

20
30
20 40 60 80 100'
Recovery of Zn (%)

Figure 7.12 ObservedmetallurgyonbulkPb/Zncleaningwiththecolumncircuit shown.


(Pb grade was between 8.5 and 12% at recoveries about 10% less than the zinc.)

prevents a strict comparison of predicted and realised metallurgy. The most


important difference is the column circuit is open, that is there is no tailings recycle
to the rougher. The open circuit was adopted in part because of the added dilution
of column tailings compared with tailings from mechanical cells. The rougher is
now pulled harder than in the old circuit to reduce the rougher tail and thereby main
tain overall circuit recovery. Probably the important changes to the column feed this
produces are a higher feed rate and a lower grade compared with the original circuit
(Table 7.5). Despite the lower feed grades, the metallurgy (Figure 7.12) is similar
to that in the original testwork (Figure 7.6).
Typical current operating conditions are shown in Table 7.6. Gas rates are
significantly lower than used in the original pilot-scale work (J ~ 2.5 cm/s). This

128
Table 7.7 Comparison of Predicted and Large Pilot Column Results at
Falconbridge ^

Feed Conditions Cu Recovery (%) Ni Recovery (%)


(t/hr) Predicted* Actual Predicted* Actual

5 72-90 72-92 0.9-2.3 0.6-1.6


10 56-80 67 0.5-1.3 0.7

*Rf over range 20% to 50%

reflects a generous capacity; by cutting gas rates the load is distributed to all three
columns. This results in a concentrate rate from the first column which is 1/2 to 2/
3 of the maximum given by C,. (Manipulating gas rate is probably a viable means
of controlling column performance in this situation.)
Simulations and plant testwork indicate 2 columns should be sufficient
most of the time. By distributing among 3 columns, however, there is greater
flexibility in responding to feed changes, particularly increases in mineral feed
rates. The third column concentrate is a candidate for recycle, with or without
regrinding; the extra capacity makes recycling a varioable option.

7.4.2 CASE STUDY 2 - Cu/Ni SEPARATION (FALCONBRIDGE)

A 0.91 m x 13 m pilot column was installed by Falconbridge to provide


both an intermediate level of scale-up and development of in-house column
operating experience.
Simulations of the pilot column were run using the kc parameters in Table
7.5b. Since Rf in the pilot column was not known, it was varied from 20 to 50%.
Column feed rate was varied from 3 t/hr (high recovery data) to 20 t/hr (low
recovery data). ResultsaresummarizedinFigure7.13andTable7.7. Thereisgood
agreement between the simulated and test column results for copper recovery
(Table 7.7), while nickel recovery was slightly overestimated. The implication of
these results is that Rf in the 0.91 m column was less than 20%. (As well, gas rate
in the 3.8 cm diameter column tests was relatively low compared to plant operation;
hence, rate constants in the 0.91 m column are likely somewhat higher than used
in the simulation.)
The circuit being implemented is a 2.1 m column as the cleaner with
existing mechanical cells as the second (cleaner scavenger) stage. It is felt that
circuit performance with mechanical cells as the second stage should be reasonably
similar to that with a column as the second stage.
The two case studies illustrate quite different conditions, Case 1 where
carrying capacity is limiting and Case 2 where capacity is limited by collection
kinetics. The scale-up procedure, while requiring refinement, accommodates both
situations. Refinement to the approach include an improved knowledge of Rf and
use of a rate constant distribution.

129
T T T
• 0.9 m pilot column data
— 0.9 m column simulation
~ 3
o
o
£
40 50 60 70 80 90 100
Recovery of Cu (%)
Figure 7.13 Copper recovery versus nickel recovery in the 0.9 m test column at
Falconbridge, showing both the simulation results and the actual results.
7.5 FLOTATION COLUMN CIRCUITS
7.5.1 SELECTED REVIEW
This review illustrates the range of circuits in use. The two basic types are
the cleaner and scavenger configurations (Figure 7.14.)
The first (Western) installation was for Mo cleaning at Les Mines Gaspe.
In that circuit, two stages of columns were used in a cleaner configuration (Cienski
and Coffin, i 981.) Gibraltar Mines chose a 3-stage scavenger configuration for Cu
cleaning, with recycle of the third column concentrate to the first column
(Redfearn, 1986.)
Flotation columns may need to be integrated into a circuit with existing
mechanical machines. This requires careful consideration of solids and especially
water balancing. Polaris have adopted three circuits depending on feed grades
(Figure 7.15.)
Harbour Lights Mining in Western Australia have commissioned a new
plant for Au-bearing sulfide flotation using columns as roughers and cleaners and
mechanical cells as scavengers (Figure 7.16). This illustrates that columns can
function as roughers, and that designs can be successfully based on data from ore
samples.
130
a)

CONC
FEED

TAD.

b>

t • CONC

FEED!

Figure 7.14 Column circuits: a) cleaning circuit, b) scavenging circuit. Dashed lines
indicate process options.

7.5.2 SIMULATION OF THE EFFECT OF SOME DESIGN AND


OPERATING VARIABLES

Column flotation is still in its infancy. There is a growing body of


experience which will accumulate in time to guide design and operation of columns
and column circuits. At present some insight is given by simulation. The effect on
performance of design options (parallel vs series units, and recycle) and operating
variables (feed percent solids and bias rate) are considered. The example used is a
Cu middlings upgrading, with the relevant data presented by Y ianatos et al. (1987b)
and del Villaretal. (1988). The points made, however, are general ones. For all
examples a 2.5 m diameter column baffled into six sections is used as the individual
column.
Parallel versus series columns (Figure 7.17): A comparison is made between
parallel and open-circuit (scavenger) series configurations. The series arrangement
shows superior metallurgy. One of the two major reasons is that columns in series
become analogous to a tanks-in-series circuit with an overall circuit transport

131
COLUMN CELL OPERATING CONFIGURATION
COLUMN CELL OPERATING CONFIGURATION
AVERAGE PLANT FEED GRADES
HIGH PLANT FEED GRADES

UUU

COLUMN CEIL OPERATING CONFIGURATION


LOW PLANT FEED GRADES

rU J

Figure 7.15 Integration of flotation columns and mechanical cells in the lead circuit
at Cominco's Polaris concentrator (Kosick et al., 1988). Large column is 0.76 m square
x 9.1 m; small column is 0.61 m diameter x 9.8 m.

character which tends towards plug flow. In the parallel arrangement the overall
circuit transport character is that of the individual columns. The second reason is
that by arranging in series the downward slurry velocity in each is higher than in the
parallel arrangement and consequently the vessel dispersion number is decreased
i.e. individual columns in the series circuit are less mixed than their counterparts in
the parallel circuit.
Recycle (Figure 7.18): The example is recycle of the 4th column concentrate to the
first stage column. The performance point moves to improved grade/recovery.
Recycle, as in circuits of mechanical flotation cells, gives improved metallurgy.
The dependence of capacity on the recycle volume must be considered in design and
operation. The recycle volume may need to be controlled to realise the expected
benefits of recycle.

132
CRUSHED ORE

CONCENTRATE FOR
FURTHER PROCESSING

Figure 7.16 Harbour Lights flotation circuit for Au-bearing sulfides. Column rougher
diameter is 2.5 m, baffled into 6 sections; column cleaner diameter is 1.2 m, baffled into
2 sections.
80

40 50 60 70 80 90 100

Recovery (%)

Figure 7.17 Comparison of series vs parallel column units (simulation result). Grade
is chalcopyrite grade. Symbols indicate 1,2,3 or 4 columns, in parallel or series
(scavenger configuration).

133
80

75

70

65

60

55

50
40 50 60 70 80 90 100
Recovery {%)

Figure 7.18 Effect of recycle of 4th column concentrate to 1st column in series circuit
(simulation result). Grade is chalcopyrite grade. Symbols in open circuit refer to 1 2 3
and 4 columns in scavenger configuration. ' '

12 3 4
Number of columns in series

Figure 7.19 Effect of feed percent solids on recovery in series circuit (simulation
result).

Feed percent solids (Figure 7.19): One advantage of columns over mechanical cells,
especially in cleaning applications, is the ability to treat feed at high solids content.
High percent solids brings the advantage of increased capacity. This is shown in
Figure 7.19 by increased recovery for column circuits (of fixed volume) with
increased feed percent solids.
Bias rate (Figure 7.20): As bias rate is increased one effect is to decrease retention
time in the collection zone. This translates into a loss of recovery (or capacity).
Maintaining JB at the minimum as dictated by grade requirements has the advantage
of maximizing capacity (and the obvious advantage of minimizing water consump
tion).

134
100

12 3 4 5
Number of columns in series

Figure 7.20 Effect of bias rate on recovery in scries circuit (simulation result).

7.6 SUMMARY OF CHAPTER

1 An approach to scale-up has been outlined, which includes the following


steps:
a) Laboratory/pilot column experiments to determine amenability, rate
constants and carrying capacity.
b) Preliminary column scale-up and circuit design.
c) Design and operation of a 0.2 to 1 m diameter pilot column to confirm
amenability data and to act as an intermediate check on scale-up. The opportunity
can also be taken for operator training and to address practical concerns related to
bubble generation and control. If the indications from Step b) are that a column
circuit will be needed, then it may be preferable to operate a pilot circuit with two
columns.
d) Full scale column design using parameters from the large pilot column
tests. (As further experience in column operation and scale-up evolves it will
become more common to scale-up directly from laboratory/pilot column data.)

2. Two case studies were presented, representing situations limited by


carrying capacity in one case and collection kinetics in the other.

3. Circuit options were reviewed and a circuit simulator was used to


examine some design and operating variables.

135
Chapter 8
Measurement and Control

Column flotation presents some novel measurement problems and control


opportunities. These are discussed here; it is not the intention to cover aspects in
common with flotation practice in general.
To provide a common frame of reference, standard control terminology
will be used where possible. To illustrate the terminology consider Figure 8.1,
based on a typical feedback control configuration (Stephanopoulos, 1984). For
example, in control of level the measured output is the level, the manipulated
variable may be tailings flowrate and a typical disturbance is feed flowrate. An
unmeasured output of interest is recovery, which the control of level may influence.
There may be a routine to estimate the recovery, this estimate being used to
manipulate tailings flowrate.
DI*turb*ftC*a

KUMurod output*
MifdptiUlad
(contrdl*d vaiUbtoa)
vutlHoa

Umn«*tiir*d output*

General Structure of Feocback Control Configuration

Figure 8.1 General terminology used for a typical feedback control configuration
(Stephanopoulos, 1984).

8.1 MEASUREMENT

This section emphasises the problem of measurement of pulp/froth inter


face position and measurement of bias flow.
137
8.1.1 PULP/FROTH INTERFACE POSITION

The pulp/froth interface position is commonly referred to as level or more


specifically as froth depth. Level measurement in a floLion column has addTZS
problems to those encountered in mechanical flotation machines. First the density
difference between the two zones in a column is not as large as it is for mechanical
mach.nesoperatingwithaconventional.drainingfroth.especiallyathigh gas rates
Second, smce the column has a small ratio ofcross-sectional area to volume the level
will respond quite rapidly to such changes as feed and gas rate.
The physical properties which have been exploited for level detection
include: pressure, density, temperature, electrical conductivity, light and ultrasonic
reflections.

8.1.1.1 Pressure - Single Measurement

This is the most common method. Pressure, measured by a variety of


techniques (e.g. transducers, manometers, bubble tubes, metritape) is sensed at
some point below the interface. Assuming predominantly static pressure is sensed
and all other input variables remain constant, then guage pressure P is linearly
dependent on the interface height above the sensor L, or froth depth H (Figure 8.2).
This linear signal is ideal for controlling level. The other input variables, however!
are subject to disturbances. The most important, in terms of their effect on P, are
slurry density, gas holdup and bubble loading (loading may change due to changes
in mineral feed rate or flotation rate). Their effect on the measured froth depth H
can be analysed mathematically. "*

Froth
Zone.f
Hf

Collection
Zone, c

Figure 8.2 Arrangement for measuring level using a single pressure sensor.

138
An example situation will be considered (Figure 8.2) and a simplified
analysis will be used (a more detailed one is given by Moys and Finch (1988b)). The
guage pressure is given by
P = (H-Hf)pcg + Hfpfg (8-la)

Therefore

Hjp = (HPcg-P)/(pc-pf)g (8-lb>

where p and pf are the average densities of the collection zone (above the pressure
sensor) and froth zones respectively, given by:

pc = Ps.n-^Wsc
Pf = Psf('V + PbSf (82b)
where psf is the slurry density in the froth and pb is the bubble density (i.e. bubble-
particle aggregate density):

p = K'KdPPP (2.14a)

The assumed conditions at the time of calibrating the pressure sensor are: H=160
cm, p =1.2 g/cm\ psf = 1 g/cm3 (i.e. negligible entrained solids), egc = 0.15 and
e = 0.70 (an average'value for the froth zone), K, = 0.5, pp = 5 g/cm3, and db =
1 mm. Two particle sizes are considered because dp has a significant effect on pb,
d = 10 |im (i.e. pb = 0.08) and dp = 50 \im (i.e. pb = 0.34).
P The problem examined is the indicated froth depth H^ upon changes in
process variables ps, and K( when the true froth depth Hf ^ is 100 cm.
For both the d = 10 ^m and 50 Jim cases, changes in ps, (1.1 to 1.3 g/cm3)
cause a marked change"in U^ (up to + 10 cm, Figure 8.3). Changes in K, have a
marginal effect in the d =10 ^im case but a more marked effect when dp = 50 u,m
(Figure 8.4).
The disturbances considered are large but not unrealistic (e.g. the range in
p represents a change in weight % solids of 11 to 28, a range frequently
encountered in cleaning circuits). It should be appreciated, however, that the effect
of the individual disturbances are additive. Consequently, errors in the froth depth
of theorderof 10cm and more(on 100 cm) must be expected with the situation being
worse with coarse (and dense) solids, because of their effect on pb. An example from
a plant installation showing a large difference between true interface level at the
same pressure setpoint (Lp im) is given in Figure 8.5; the difference corre
sponded to an observed change in froth appearance between trial 1 and trial 2.
The literature reveals conflicting views on the effect of froth depth on
metallurgy. It may be partly because level is not sufficiently well measured, and
thus controlled, for the effect of Hf to be revealed. The impact of this error in Hf
however, must be kept in perspective. Operations with deep froths (~ 100 cm) and

139
90
1.1 1.2 1.3
Slurry density, P,t (g/cm1)

Figure 8.3 Simulated effect of sluny density on froth depth indicated by the
pressure sensor H^, for conditions in text.

no

100

0.4 0.5 0.6

Fraction of a monolayer, K,

Figure 8.4 Simulated effect of bubble loading (K,) on H,,,, for conditions in text.

fine particles (dg0 < 30 urn) are unlikely to experience serious metallurgical vari
ations attributable to errors in H^. Operations with shallow froths (< 50 cm) and
coarse, dense particles may find the error in H^ produces unacceptable variations
in metallurgy. Methods of improved level measurement are warranted on two
counts: first, optimizing control will require tighter control of level; and second,
techniques being considered for bias control require knowledge of the true interface
position.

140
o Trial one
60
a Trial two

§ 40

20

"20 30 40 SO

LPsei-point (0-100% scale)

Figure 8.5 Example of change in L^ upon change in froth character at the same
L . The froth in trial two was visibly more loaded. L is distance from the pressure
-pinpoint _
sensor to the interface.

8.1.1.2 Multiple pressure sensors

The problem of level measurement with a single pressure sensor is the


dependence on pc and pf (equation (8.1b)) which vary with operating conditions in
a manner that is difficult to predict. Attempting to allow for change in pc and pf by
including, for example, an on-line measurement of feed percent solids, may actually
compound the error. By arrays of pressure sensors, however, in effect pc and pf can
be estimated continuously and hence Hf estimated with more accuracy.
An arrangement is shown in Figure 8.6 with two sensors in the collection
zone and one in the froth zone. The pressure at each sensor is

P| = (8.3a)

P2 = (8.3b)

P3 = H3p,g (8.3c)

From equation (8.3a) Hf is given by (see equation (8.1b)


141
Hf = (H1Pcg-PI)/(pc-pf)g (8.3d)

and the densities pc, pf are given by, respectively

(8.3e)

Pf = (8.30

Thus, from sensors P, and P2, pc is estimated (equation 8.3e) and from sensor P ,
pf is estimated (equation 8.40. These values are substituted into equation (8.3a) to
solve for Hf (this estimate of Hf will be called H )

H = H3[H,(P,-P2)- P.CH.-H,)]
fP(3) (8.3g)
[H3(P,-P2)- P3(H,-H2)]

Thus the pressure signals from the three sensor arrangement can be combined to
give froth depth independent of variations in pf and pc. Other arrays of pressure
sensors are possible, for example two sensors in the froth.

H3

H, H,

H,

Figure 8.6 Possible arrangement of three pressure sensors for measuring level.
142
The improved fit to Hf ^ of the three sensor result (H ) over that of the
one sensor (UmJ is shown in Figure 8.7 for laboratory scale trials (Dirsus 1988).
Someohhe limitations to this method are: the need for Hf to be between
P and P (which may argue against fixing the sensors in the wall in favour of
mounting them on a probe so that Hf can be varied over a wider range); the
assumption that the collection and froth zone densities (pc, pf) are uniform up to the
interface (i.e. there is no density profile) and, the assumption that the dynamic
component of the pressure is the same in the froth and collection zones.

60 120 180 240

Time (s)

Figure 8.7 Improved measurement of froth depth (H( ^J using 3 pressure sensor
arrangement (Hm3)) compared with single sensor arrangement (HW)). Laboratory tests
(column dia. 3.7 cm, height 250 cm) were conducted using silica slurry with changes in
feed percent solids as the disturbance.

8.1.1.3 Float

In this method an object of density intermediate between the froth and


collection zone is used to float at the interface. By means of a rigid vertical arm the
movement of the interface can be sensed (e.g. Hoffert, 1987). (In testwork requiring
knowledge of Hp lowering a float on a string and determining when the string
becomes slack can be used.)
The success of the float method relies on the difference in pc and pr The
impact of variables on pc and pf can be analysed using equations (8.2a) and (8.2b);
high bubble loading with coarse, dense particles and relatively low slurry densities
is the worst combination (Figure 8.8). It should, however, always be possible to
predetermine a float density intermediate between pc and pf
When a float is used in a mechanical cell it can be located in a tailings box
to exclude froth (i.e. pf becomes 0). This is not applicable in a column. Solids
buildup on the float has caused problems in conventional cells and this may be a

143
or

a-

Figure 8.8 Effect of bubble loading (K,) on pc, p^ conditions are the same as for
Figure o.J.

greater problem in a column where the (pc - pf) is tighter. The float geometry may
be important to minimise solids buildup.

8.1.1.4 Temperature

This is a proposed method initiated from the observation that the wash
water in most operations is cooler than the feed pulp. Provided there is efficient
rejection of entrained water at the interface, a marked change in temperature occurs
at the interface, that can be detected by an array of temperature sensors (Figure 8.9).

Figure 8.9 Example of temperature profile across the interface in a 0.76 m square
column. The profile was measured by a series of thermistors mounted on a probe
Reduced temperature ^ is given by equation (8.4).
144
To compensate for changes in feed pulp temperature (Tp) and wash water
temperature (T^a dimensionless (reduced or normalised) temperature was
defined <j>;
V ' (8.4)
T -T w

whereT.isthetemperatureatadistancez.fromthetopofthecolumn. The interface


can be located by a large change in<p;, where

A model to estimate H^ was suggested by Moys and Finch (1988b):

(8-5)
Hfr =

where m is a user selected constant.

The method has proved successful in plant trials (Figure 8.10). Use of the
method clearly depends on there being a large enough difference in temperature
between the wash water and the feed pulp (the wash water could be warmer or
cooler). Temperature can measured using either thermistors or thermocouples, the

eo
o Trial 1

A Tritl 2

E
- 40

20

Localion of ••fl«ori

8, ^, S,

20 40 60 80 100

"I,™. <cm>

Figure 8.10 Correlation between measured froth depth H^ and true froth depth Hf ^
using the temperature based technique. Accuracy increases with the number of sensors.

145
former bemg more precse but less rugged. With thermistors a temperature differ
ence as small as 4°C is sufficient, with a Type E thermocouple minimum

suffi^, Hff y grinding ^ Pri°r n°tati°n St3geS- N^rthdess whether


asufficentdifferenceintemperatureisavailablewillbeacase-by-casedetennina-
tion Dehberate heating of the wash water to produce a temperature difference is
unikely to be economic unless waste heat is available or other benefits such as
improved cleaning efficiency are also achieved.
In the absence of a significant temperature difference between feed and
wash water it may still be possible to detect the level using thermistors because of
the expected large difference in thermal coductivity between pulp and froth This
difference is expected because of the different relative amounts of solid gas and
liquid in the two zones.

8.1.1.5 Electrical conductivity

A difference in electrical conductivity between the froth and collection


zones occurs because of differences in gas holdup, salt content (wash water
probably has a lower content than the pulp water) and temperature. On the
basis of differences in gas holdup alone the relative collection zone to froth zone
conductivity is approximately given by (Yianatos et al., 1985)

Kc= (1-8.) (2.315 £ef)


Kf (l+0.55£gc)(l-egf) (8'7)

For typical values of e^ (0.15) and £gf (0.7) the relative conductivity is about 3.3.
Plant measurements indicate a range from 2 to 15.
The interface position could be detected, therefore, by an array of
electrical conductance sensors. This is demonstrated in Figure 8.11 where the sharp
change in profile clearly corresponds to the visually located interface. There are a
number of possible sensor arrays. One design being investigated is a series of ring
electrodes mounted on a circular cross-section probe. A commercial device, based
on electrodes at the tip of several probes of different length, is already in use in a
notation column (Nicol et al., 1988). A design giving the profile is to be preferred
over devices giving a single signal. An example of the latter is a long pair of
electrodes passing through the interface (Moys and Finch, 1988a). The signal from
such a design will only remain proportional to froth depth provided the froth zone
and collection zone conductivity does not change with time. In most cases this
provision will not be met. By profiling, changes in froth and collection zone con
ductivity do not matter.
One development has been the design of a probe for calibration purposes,
Figure 8.12. An example of re-calibration of a single pressure sensor is shown
(again indicating the error that can be encountered using a single pressure sensor).

146
|82
If 1.1 2
Is
IB Froth depth estimated by pressure
prior to re-calibration (m)
Distance from top of the cotumn (cm)
to
§

I
Is
I
O
o g

««-s !■«■
B* 3(D
.. . o.

3 r*

I I 1 L.
8.1.2 BIAS

8.1.2.1 A problem of definition

Throughout the book care has been taken to define the bias rate as the net
downward flow of water through the froth, i.e. JB (as a superficial velocity) or its
equivalent the net difference of water flow between the tailings and feed It is
important to appreciate that the definition often used in control is based on a net
difference of slurry flowrate in tailings and feed. In superficial velocity terms (i e
dividing by Ac) this difference JT F is

Jtf = JT - Jf (8-8)
The flowrates JT and JF are measured by magnetic flowmeters, for example. (A
further confusion is the use of a bias ratio J^ this will not be considered here.)

The relationship between J and J can be derived from a volumetric


balance (Figure 8.13a).

Volumetric balance (dividing throughout by A ):

overall: Jf+Jw=Jc + Jt (8-9)


cleaning zone: Jw + J^ = JB + Jc (8-, 0)

collection zone: JF + JB = JCs + JT (8.11)

where Jcsis ^ volumetric flowrate of solids reporting to the concentrate per unit
column cross sectional area. Substituting equation (8.8) into equation (8.11) then,

Jb=Jcs+Jt-f (8.12)
Equation (8.12) indicates that JB is greater than JT p by the amount J ; J can be
of the order of JB for high feed percent solids at high solids recovery.

A numerical example is shown in Figure 8.14 The calculations refer to dc


= lm, JF = 1.06 cm/s (QF = 500 L/min), 80% solids recovery (a figure approached
in cleaning circuits) and varying volumetric percent solids in the feed (i.e. varying
Jcs )•Two set P°ints are considered, JB = 0.1 cm/s and JT F = 0.1 cm/s. Above about
12% solids by volume, JCs becomes greater than 0.1 cm/s, and thus withJ set,
JT F becomes negative; or with JT F set Jfi becomes greater than twice JT p. B
The control of bias by JT F has some ramifications: JB variesT\vith feed
percent solids; JB is always larger than it need be (JB < 0.1 cm/s is usually sufficient);
at high feed percent solids this large JB may create undue demand for water and loss
of capacity in the collection zone.

148
J

Cleaning
zone

g/3

Collection
zone

Figure 8.13 Volumetric flowrates in and around the column. Flowrates are given as
superficial rates, i.e. Q/Ac where i = F,C,T,W,B and Cs, where Cs refers to solids
reporting to the concentrate. Flowrates are slurry flowrates in case of F,C and T.

Jcs lcm/11)

0.045 0.09 0.135

0.20

0.15

•i 0.10

0.0S

5 10 IS
Food parcant solids by votum*

Figure 8.14 Example of effect on JB if JFT is set, and effect on JF T if JB is set, for
conditions in text.

Compared with JT F, JB is the preferred variable to control; the question


is how to measure it.

8.1.2.2 Mass balancing for measurement of JB

In principle by measuring flowrates and pulp densities around the column,


followed by mass balancing, the net flow of water across the interface can be
149
estimated There are several difficulties. First, concentrate flowrates and densities
XlTpossible to obtain- ™sthrows * phasis
hi on ^ SS
Wlth no measurements on the concentrate
Jb = [QT0-<U-QF0-<l>lF)]/Ac (8.13)
where <psT, <psF are volume fraction solids in the tailings and feed slurries A
problem with this approach is that it is susceptible to error propagation As an
example, consider a 2 m diameter column with process flow and density measure
ments on tailings and feed streams of: QT = 1.90 m3/min, d> T = 0 09 Q =199
m /mm and <j>sF = 0.18. Assume for this case a relative standard deviatkTn of 3%
tor each of the process measurements (e.g. standard deviation of Q is 3% of 1 9
- 0.057 m /tain). Then, from equation (8.13), JB = 0.052 cm/s and! using Taylor
series expansion for estimation of the standard deviation of J_, the relative standard
deviation of JB is 74%; the 95% confidence interval on J_ is 0.13 to -0 02 Clearly
there is the potential for large error in JB with this method. It relies on excellent'
calibration and high precision of slurry flow and density measurements An energy
balance could be used to augment the balance data if there exists a temperature
difference between feed and wash water.
A further problem is that the system is dynamic and estimates from mass
balance calculations assume steady state. For well controlled tests at the laboratory
or pilot scale mass balancing to estimate JB is quite adequate.

8.1.2.3 Property profiles for measurement of J


B

It was seen in Figure 8.9 that a sharp change in temperature could be used
to locate the interface. The Figure also reveals a temperature profile through the
froth zone, a profile which depends on operating variables including J (Moys and
Finch. 1988a). The possibility exists, in principle, of manipulating J to control to
a set-point profile corresponding to efficient froth cleaning.
This technique, therefore, does not measure JB but measures a profile
which is dependent on JB. An example is shown in Figure 8.15, where it took about
10 minutes to go from profile 1 to profile 2 and from profile 2 to profile 3 upon step
changes in Jw as indicated (the correlation is with Jw since J is unknown). That
the profile is related to the metallurgy is revealed by the % Fe contaminant in the
overflow product. Note that the temperature profiles are essentially vertical through
the froth zone, suggesting the zone is well mixed in this case.
Properties other than temperature may be usuable. Experience with a two
phase (gas-water) system has revealed that conductance profiles through the froth
are sensitive to JB. There is a possibility that pressure profiles can be interpreted to
yield information on Jn.
The attempt to measure JB is spurred by the desire to control overflow
grade. This presumes grade can be infered from JB. If on-stream analysis is
available then grade can be controlled directly. However, as column circuits grow

150
Location of
temp, sensors

N
e

"5
o

0.2 0.4 0.6 OS


Reduced temperature, o,

Figure 8.15 Example of shift in temperature profile upon changing Jw and resulting
effect on overflow Fe contaminant grade. (For other operating conditions sec Figure 8.9).

in size it is unlikely that each unit will have on-stream analysis, making inferential
methods based on measuring bias worth evaluating.
To date on-line measurement of JB remains a challenge.

8.1.3 GAS RATE AND GAS HOLDUP

8.1.3.1 Gas rate

The reason for this section is for two reminders. First, standard condtitions
(usually STP) should be used. STP actually gives the flowrate through the top of
the column. Flowrate at any level in the column can be calculated knowing the
pressure at that level. For example, a 12 m column with slurry exerts a pressure at
the sparger level of about 2 x atmospheric, thus the volumetric flowrate at the
sparger is about one-half that at the top of the column (see Appendix B).
Second, if a rotameter is used for measuring gas flowrate then the
regulating valve should be downstream of the rotameter in order to provide a
measure of the STP gas flow by measuring the line pressure and correcting the
rotometer gas flow for pressure. If the regulating valve is placed on the upstream
side of the rotameter (and the rotameter is calibrated in this manner), then the
pressure at the rotameter, and therefore the true flowrate, is unknown.

8.1.3.2 Gas Holdup

Differential pressure is the common technique for measuring gas holdup


in commercial columns. In order to use it the slurry density must also be measured
(equation (2.1)). The recommended solution at the moment is to measure AP over
151
a short (eg. 2 to 3 m) section near the bottom of the column so that the slurry density
of the underflow approximates the required p The lower pressure sensor should
probably be about 1 column diameter above the spargers to avoid disturbance from
the turbulence of gas injection.

8.2 CONTROL

A convenient division is to consider stabilizing control and optimizing


control. Stabilizing control, which must be established first, demands as a minimum
control of level, with bias control as a useful addition. Optimizing control requires
control of grade and recovery (frequently unmeasured variables) according to
economic criteria.
Most of this section will concern control of individual columns; the last
section (8.2.4) offers some general remarks on column circuits. Whether consid
ering stabilizing or optimizing control of an individual unit a useful first step is to
develop a process matrix.

8.2.1 THE PROCESS MATRIX

The process matrix is a summary of the interactions between controlled


and manipulated variables gained from experience and knowledge of the process.
Table 8.1 is the authors' version of the process matrix for flotation columns. The
Table shows the response of the controlled variable to an increase in the manipulated
variable. Each manipulated variable is considered in turn with no simultaneous
change in any other manipulated variable. The Table shows the direction of the
response (+ means controlled variable increases) and speed of response (S means
slow, etc.). Ynchausti et al. (1988) have also presented a process matrix for
columns.
It is worth reviewing the interactions, especially since a few may not be
that obvious; indeed others' experience may contradict some of the responses
identified in the Table.
Wash water rate: The effect on level tends to be slow, since it takes time for the water
to work its way through the froth bed. Control of level by wash water is practised
(Coffin and Miszczak, 1982). It has a slow effect on bias rate for similar reasons.
In fact to affect bias rate it is necessary to control level constant, for example by ma
nipulating tailings rate. Wash water is shown to have a fast effect on grade. The
effect, however, may be + or -. Remember, if wash water rate is already high a
further increase may cause excessive mixing in the froth and unselecti ve carry-over.
The effect of an increase in wash water and increase in bias will be to reduce
collection zone retention time and hence reduce recovery. However, the same effect
on mixing noted above may cause an increase in froth removal rate and hence
increased recovery. This may be a temporary phenomenon, or may indicate an
increased froth zone recovery.

152
Tailings rate: This has a fast negative effect on level (or positive effect on froth
depth). It also increases bias rate.
Gas rate- Gas rate has a fast effect on level. Initially level will nse as the volume
of slurry displaced by gas at the bottom of the column increases. However, the new
steady state level, if uncontrolled, may be lower, as the increased water (slurry)
transported across the interface into the froth zone may offset the effect of increased
gas holdup Gas flowrate generally has a positive effect on gas holdup but it may
be a zero or even negative effect if gas rate is already high. This is related to
transition from bubbly to churn-turbulent flow. The transition is characterised by
unstable gas holdup values, which may translate into a decrease in recovery, and is
probably related to the phenomenon described as' impending holdup' (Amelunxen.
1985). Manipulating gas rate to control gas holdup is a loop in use.

Table 8.1 Process Matrix for a Flotation Column-Response of Controlled


Variables (Measured and Unmeasured) to Increase in Manipulated Variables

+ controlled variable increases as manipulated variable is increased


- controlled variable decreases as manipulated variable is increased
S slow response
F fast response
M moderately fast response
blank means no expected response
( ) indicates steady state response if possibly different from dynamic response.

153
Frotherdosage: This, along w.th feed rate, is less likely to be used as a manipulated
vanab e. There tends to be a slow response to frother dosage principally because
of the time it takes for the frother to distribute throughout the column. Changing
frother dosage ,s the way to move from one regime of gas holdup versus gas rale to
another, if this is desired. There may be a rapid increase in water content of the
concentrate if frother is added into the wash water. This can make the froth flow
more readily, and thus increase solids removal rate. This effect will probably be
most evident in columns where the wash water inlet is below the overflow lip and
a conventional, drained froth is formed. The effect may be temporary
Feed rate: This does have the virtue of a fast effect on level which could be
exploitable if surge capacity ahead of the column is available. In at least one
operation a control loop to maintain a stable feed flowrate is used (Aravena, 1987).
The process matrix shows that there is interdependence, i.e. one manipu
lated variable affects more than a single controlled variable. This has to be
recognized in developing a control stralegy.

a)

.CONC

FEED
-—©

AIR

I LEVEL
SCT-POtMT
i
TAILS

Figure 8.16 Stabilizing control schemes in use. FDC - Flow difference controller
(FRC - flow ratio controller could be substituted); FT - flow signal transmitter; FI - flow
indicator; LT - level signal transmitter; LC - level controller
a) control of level only (e.g. Mauro and Grundy, 1984)
b) control of level and bias (e.g. Coffin and Miszczak, 1982)

8.2.2 STABILIZING CONTROL

Figure 8.16 shows two stabilizing control schemes which have been
successfully implemented. The scheme in (a) is about the simplest (and cheapest)
possible; level is controlled by manipulating tailings rate but there is no direct
control over bias, and wash water addition is manual (Mauro and Grundy, 1984).
This strategy will work provided gas rate is kept below a certain level. For.example,
keeping Jg < 1.5 cm/s ensures that the volume of water carried into the froth is not

154
sufficient to drive the system into negative bias. Less critical, perhaps, is also the
need to avoid excessive wash water which may mix the froth. Maintaining a deep
froth (~ 100 cm) would tend to dampen the detrimental effects of excessive gas and
bias rates. _ „ _ ,
The scheme in (b), developed by Column Flotation Co. of Canada, ma
nipulates wash water to control level, and tailings rate to control the flow difference
between tailings and feed i.e. JTF (Coffin and Miszczak, 1982). (A slightly
modified approach using bias ratio was used by Amelinxen (1985).) Thus, unlike
(a) this method does make an attempt to control bias (although not JB). According
to the process matrix, this method of level control will probably exhibit a slow
response.
The two schemes were compared by Feeley et al. (1987) in a 1.1 m
diameter column on Cu-Ni separation (with scheme (b) using bias ratio). Both gave
similar metallurgy but scheme (a) was favoured as control was simpler and involved
less instrumentation.
From the process matrix the suggested stabilizing control strategy is given
in Figure 8.17. In (a) the bias JB is from mass balancing, in (b) it is assumed JB can
be successfully inferred.
The stabilizing control strategy should be kept simple. The scheme in
Figure 8.16(a), provided excessive Jg and JB are avoided, may prove adequate in
most cases, and certainly can be recommended for start-up testwork. The suggested
b)
WASH WATER WASH WATER

CONC CONC

AIR AIR

(C 5CT-PG(KT

TAILS TAILS

Figure 8.17 Stabilizing control schemes suggested by the process matrix (a) bias
from mass balance (b) bias inferred.

method of level control is by manipulation of the tailings flow rate which will
probably involve a valve (as opposed to a variable speed pump). This valve is the
only major source of wear in the column. (For laboratory/pilot-scale control work,
a variable speed pump is essential to manipulate tailings rate; at the low flow rates
encountered at this scale valves of any type tend to plug giving erratic, surging
flows.) In some situations, level control may be achieved by directing the tailings

155
through a vertical rise pjpe ^ manipulating ^ leve, fn)m whkh the ^

discharge. Another possibility is manipulating feed rate from a surge tank


Whatever the choice, simplicity and reliability are the criteria.

8.2.3 OPTIMIZING CONTROL

The manipulated variables that have been tested to control metallurgy are
gas rate/gas holdup, bias/wash water rate and level. Their use will be illustrated by
examples from practice with observations offered based on current understanding.
Each installation will need to conduct its own investigations.

32
u

31 \

a>
30
a
a
o
u
29 §

3.0 3.5 4.0 4.5


Superficial gas velocity, Jg(cm/s)

Figure 8.18 Grade/recovery as a function of gas rate at Magna Copper Co. (Clingan
and McGregor, 1987). (The gas rates are higher than usual, suggesting relatively large
bubbles.)

8.2.3.1 Gas rate/gas holdup

Manipulating gas rate appears to be the most common method of control


ling grade and recovery. An example is shown in Figure 8.18 from Magma Copper
Co. (Clingan and McGregor, 1987). Increasing gas rate decreased grade and
increased recovery. This result reflects one or more of the following: (i) increased
collection rates, (ii) increased froth zone recovery, or (iii) increased entrainment.
Another approach is to relate recovery to ihe volume of gas per unit of mineral in
the feed (Coffin and Misczczak, 1982).
There is an upper limit to gas rate, which may have a variety of origins
(Table 8.2). This maximum will be determined by metallurgical considerations or

156
operatingconstraints. Operating constraints include breakthrough of large bubbles
at the froth surface (commonly referred to as "burbing") and erratic signals from
level and gas holdup sensors. .
A control strategy based on gas holdup (with gas rate as the manipulated
variable) seems attractive, in that gas holdup is a function of bubble size and gas rate
both of which influence grade and recovery. Gas holdup, however, is both difficult
to measure (because of a need to simultaneously measure p^) and is a function of
several variables. Consider the following possibility: feed flowrate increases, level
is held constant by increasing tailings rate, which in turn increases gas holdup
calling for a decrease in gas rate. A decreased gas rate does not seem the desired
end response to an increase in feed rate. In reality gas holdup may prove sufficiently
insensitive to such changes over a normal range. To date the one published attempt
to relate metallurgy to gas holdup is not convincing (Dobby et al., 1985).

8.2.3.2 Bias/wash water rate

There is limited data relating performance to bias rate, as opposed to wash


water rate. One example was given in Chapter 5; with bias rate over a range JB =
0 to 0.2 cm/s there was little effect on grade (Figure 5.19). With increasing bias rate
recovery is expected to drop due to decreased retention time in the collection zone;
grade may eventually drop due to increased mixing in the froth zone.
These same general observations will apply to wash water rate also. Figure
8.19 shows a result from Cominco's Polaris concentrator (Kosick et al., 1988); with
increasing wash water rate grade increased and recovery decreased. In at least one
case increasing wash water increased recovery (Coffin, 1981). This probably
reflects a loaded froth and additional wash water was needed to discharge the
concentrate. Similar reasoning was used by Amelunxen et al.( 1988) in developing
a recovery control scheme.

Table 8.2 Maximum Gas Rates Reported in a Variety of Installations

Location Duty Column Dia Maximum J Comments


(cm) (cm/s)

157
0.08 0.16 0.24 0.32 0.40
Superficial wash water velocity,Jw(cm/s)

Figure 8.19 Effect of wash water rate on grade and recovery at Cominco's Polaris
concentrator (Kosick et a!., 1988).

8.2.3.3 Level, froth depth

There is no clear evidence on the importance of level. In one case no effect


of level was reported (Clingan and McGregor, 1987), in another level was included
in the control strategy (Amelunxen et al., 1988). Part of the problem is inaccurate
level measurement. Although the evidence is mixed it is reasonable to expect that
too low a level will cause loss of retention time in the collection zone (and hence loss
in recovery) and too high a level may lead to inadequate cleaning and/or selectivity
and loss of grade.

8.2.4 CIRCUIT CONTROL

Most control studies have considered single units. Optimizing control of


column circuits is not necessarily achieved by optimizing the individual units.
Two types of circuit were shown in Figure 7.14. They can be designated
a cleaning circuit (Figure 7.14a) where the concentrate (or at least overflow
product) is reprocessed, and a scavenging circuit (Figure 7.141b) where the tailings
(underflow) is reprocessed. A number of possible recycle options were indicated,
including no recycle, i.e. open circuit.
The scavenging circuit arrangement reflects the common situation where
controlling grade is relatively less important than controlling recovery. This type
of circuit was adopted at Gibraltar (Amelunxen, 1985). Currently the circuit there
is 3 stages with the third stage concentrate being recycled to the first stage. Mt. Isa
Mines adopted three stage open-circuit scavenging for upgrading of bulk Pb/Zn
158 ,
(see Section7.4.1).Thefollowing observations referto this three-stage scavenging

Optimizing control of the scavenger circuit will likely emphasise the third
column where the final grade/recovery operating point is made. At Gibraltar Mines
the third column is operated at high gas rateand slightly negative bias to ensure high
Cu recovery. At Mt Isa Mines the control effort emphasises the third column where
gas rate is being investigated as the manipulated variable.
Part of the optimizing control strategy may be to re-route the third column
concentrate. For example, if feed rate or feed grade decreases, concentrate grade
from the third column will decrease, and could be recycled either to the first or
second column. This flexibility is worth building-in at the design/construction
stage. The penalties associated with recycle should be appreciated, however.
Recycle will reduce flotation retention time which must have been allowed for. A
fluctuating recycle volume may be more disruptive to control than beneficial;
controlling the recycle volume may be a required control strategy. There is also the
unresolved question on the effect of recycling to a column which is at its carrying
capacity.
At this stage of still limited experience with flotation columns, and
particularly circuits of columns, little more than these generalizations can be
offered. As a final point, if columns are to be integrated into a circuit with
mechanical cells interaction between the two flotation devices needs careful
consideration because of their different metallurgical and operating characteristics.
One example of integration is at Polaris where the lead circuit configuration
depends on the head grade (Figure 7.15).

8.3 SUMMARY OF CHAPTER

1. The common method of measuring level (or froth depth) based on pressure
from a single sensor is subject to error, as the zone densities are variable. A more
accurate measurement of level is possible with either multiple pressure, tempera
ture or electrical conductivity sensors.

2. Bias can be defined in different ways; the bias required is the net flow of
water through the froth. Calculating bias from on-line mass balance is subject to
large error. Bias may be inferred by temperature or electrical conductivity profiles
through the froth.

3. A process matrix has been presented.

4. Stabilizing control can be achieved by manipulating tailings rate to control


level, provided upper limits for gas and wash water are respected.

159
FDCF-l
5. The stategy for optimizing control has been to manipulate gas and wash
r^Sr^- A l ^ £
6. Control of column circuits needs to be critically examined. A possibility
is to emphasise control of the last column in a series, including provision for
concentrate recycle.

160
APPENDIX A

Approximate operating conditions at selected column installations


Data is derived from discussions with plant personnel, results of a survey,
and journal publications.

Cominco Polaris, S=0.76 rubber 2.4 0.3 0.6 20 40/74

Canada Hfc=9
Coarse Pb 1 section

Magma Copper, dc= 1.8 fabric 3.5 0.2 0.8 8 N.A./N.A.


U.S.A. Hf=12
1st Cu Clnr 1 section

Harbour Lights, dc=2.5 fabric 1.7 N.A. 1.2 II 35/60

Australia HfC=12
Pyrite Rghr 6 sections

Cominco Sullivan, dc=2.4 USBM 1.5 0.3 0.9 11 45/50

Canada Hfc=13 type

Zn 1st Clnr 6 sections

Falconbridge, d=0.9 fabric or 1.7 0.4 1.0 25 30/50

Canada Hfc=13 USBM


Cu/Ni Separ. 1 section type

S = side, if square or rectangular


161
APPENDIX B

Correction for pressure (column height) on gas rate, bubble size and gas holdup
The governing equation is
PQ = constant

where P is absolute pressure at any level in the column and Q is the volume flow
rate of gas at that level. Consequently, PJg is constant. Therefore, at any level in
the column with absolute pressure P(inkPa) the gas velocity
Jg(aiP>isgivenbv
j = j p /p
g(al P) g(at atm) aim '

where Paim is 101.33 kPa at STP.


In the text Jg(alalm) is referred to as simply J .

The corresponding bubble diameter is given by

"b<ai am,) (Pa

Table B1 is an example calculation to illustrate the effect of pressure (i.e. distance


from top of the column) on gas velocity, bubble diameter and gas holdup. The data
in the Table is generated by assuming Jg(at m) = 1.5 cm/s and the corresponding
bubble diameter dJ^Matm) is 1.0 mm. Gas holdup for each case has been calculated
using equation (2.6) (see also Appendix C). Gas holdup decreases from top to
bottom of the column; this implies that care must be taken in comparing gas holdup
versus gas rate from different plant operations. (Similar results to the simulation
values of Table Bl have been observed in large diameter columns.)

Table Bl Calculated effect of column height on J , db and e for a gas/water


system (water velocity is 1 cm/s downward) % s

* approximate distance corresponding to pressure

162
APPENDIX C

Example calculations of gas holdup and bubble diameter

Case 1 - GAS/WATER SYSTEM

Consider a 1 m diameter column where guage pressure measurements PA and PB


are indicated at points 11 m and 8 m, respectively, below the top of the column (as
in Figure 2.2a). The column is being fed with water only (with an appropriate
concentration of frother) in a conventional countercurrent mode. Assume the
following conditions are measured:
P = 92 32 kPa and PD = 68.60 kPa; Qo (at STP) = 1600 L/min; and QT = 600 L/
A B g
min.

1. Gas holdup €: AL = 3.0m and AP = 23.72 kPa.


Therefore, from equation (2.1)

23.72
e = 1 =0.194 (or 19.4%)
8 (10)(981)(30)

2. Gas velocity J (midway between pressure points):


Since 1 atmosphere = 101.33 kPa, Qg midway between the pressure
measurement points is given by
101.33
Q =16OO[ 1 = 1600(0.557) = 892 L/min
8 101.33 + 0.5(68.6 + 92.32)

Then J = 892/Ac and, after converting units to cm and seconds,


Jg =L89 cm/s (Jg(atatm) = 3.4 cm/s)

3. Slurry (or liquid) velocity J.,: Js| is given by Qj/\- Therefore,

Js| = (600 L/min)/Ac


S = 1.27 cm/s
1.89 1.27
4. From equation (2.5) U. = + ; = 11.3 cm/s
Ss 0.194 (1-0.194)

5. The bubble diameter that gives USg =11.3 cm/s in equation (2.6) is
calculated by iteration, with the following result:

db = 1.19 mm
(and Ub = 14.4 cm/s; Reb =171; m = 2.67; Re^ = 109)

163
This is the estimated bubble diameter at the mid point of the holdup
measurement. At 1 m below the top of the column (approximately at the top of the
collection zone) and neglecting mass contributed by the froth, the pressure is 0 557
x the pressure at the point midway between the pressure taps (see Step 2).

119
db.op= = 1-45 mm.
p (0.557"3)

Case 2 - GAS/SLURRY SYSTEM

Assume a similar column configuration as in Case 1, with the following


conditions:
PA = 107.01 kPa and PB = 75.63 kPa; Qg(at STP) = 1600 L/min; Q = 600 L/min
slurry; and psl in tailings = 1.25 (assuming p = 4.0, this gives volume percent solids
= 8.3 %). p
It will be assumed that ps| in the column between the pressure taps is the
same as the ps, of the tailings (true for fine sized particles). As well, it is assumed
that the gas bubbles are only lightly loaded with solids, since measurement is made
at the bottom end of the column; therefore pb = 0. (If the assumptions regarding p
and pb are considered to be invalid, then the computation of db is somewhat more
complicated; this is examined elsewhere (Dobby et al., 1988 and Yianatos et al
1988c).)

1. Gas holdup 6g: AL = 3.0m and AP = 31.38kPa


Therefore, from equation (2.1)

31.38
e = 1 =0.147 (or 14.7%)
g (1.25)(9.81)(3.0)

2. Gas velocity Jg (midway between pressure points): Q midway between


the pressure measurement points is given by
101.33

Then Jg = 842/Ac = 1.79 cm/s

3. Slurry velocity Js|: This is approximated by using the tailings velocity,


Js| = (600 L/min)/Ac
= 1.27 cm/s

4. From equation (2.5) USg = Jj^- + ±?L^- = 13.67 cm/s

164
5. Slurry viscosity can be estimated from one of many correlations. The one
used here is that of Roscoe (1952):

= |lw(l -(JTr2 = 0.01 (1 -0.083)-25 = 0.0124 poise

6. The bubble diameter that gives USg = 13.67 cm/s in equation (2.6) js
calculated by iteration, with the following results:

d. = 138 mm
(and ub = 16.1 cm/s; Reb = 223; m = 2.61; RebS = 162)

This is the estimated bubble diameter at the mid point of the holdup
measurement zone. At 1 m below the top of the column (approximately at the top
of the collection zone) and neglecting mass contributed by the froth, the pressure
is 0.526 x the pressure at the point midway between the pressure taps (see Step
2).

1.38
Then d = = 1-71 mm.
lhe" a»«>P (0.526"3)

Table Cl gives results of Case 2 type calculations on a pilot column (dc = 0.91 m)
operating at Falconbridge's Strathcona concentrator. Note the general trend of
increasing gas holdup and bubble size with increasing gas rate, in agreement with
the observations made in Chapter 2.

Table Cl Gas holdup and bubble size versus gas rate measured on a 0.91 m
diameter column used for Cu-Ni separation (Huls, 1988)

* measured at mid point of a 13 m high column

165
References

Al Taweel, A.M. (1987) in Column Flotation Cell Symposium Proceedings


(Feasby, ed.), Trail, B.C., Canada, published by CANMET, Ottawa, June 1988
Amelunxen, R.L. (1985) The mechanics of operation of column flotation machines
Proc. 17th Annual Meeting of the Canadian Mineral Processors, CI M Ottawa'
January, p. 13-31 ' "

Amelunxen, R., Llerena, R., Dunstan, P. and Huls, B. (1988) Mechanics of column
flotation operation, in Column Flotation '88 (K.V.S. Sastry, ed.), SME Annual
Meeting Phoenix, Arizona, p. 149-156

Anfruns, J.F. and Kitchener, J.A. (1977) Rate of capture of small particles in
flotation, Trans. I.M.M., 86, p.C9-15

Aoyama, Y. Ogushi, K., Koide, K. Kubota, H. (1968) Liquid mixing in concurrent


bubble columns, J. of Chem. Eng. of Japan, 1(2), p. 158-163
Apian, F.F. (1976) Coal flotation, in "Flotation - A. M. Gaudin Memorial Volume"
(M. C. Fuerstenau, ed.), AIME, N.Y., p. 1235-1264

Aravena, J.J. (1987) Column flotation applications at Chuquicamata's molybdenite


flotation plant, in Proceedings Copper '87 (A.L. Mular, G. Gonzalez and C.
Barahona, eds.), published by the University of Chile, p. 155-170
Argo, W.B. and Cova, D.R. (1965) Longitudal mixing in gas-sparged tubular
vessels, I&EC Process Design and Devel., 4(4), p.352-359

Baird, M.H.I, and Rice, R.G. (1975), Axial dispersion in large unbaffled columns
Chem.Eng.J.,9,p.l71-174

Bascur, O.A. and Herbst, J.A. (1982) Dynamic modelling of flotation: A view
toward automatic control, Proc. 14th Int. Mineral Processing Congr., Toronto,
Canada, Paper 3-11

Bhaga, D. and Weber, M.E. (1972) Holdup in vertical two and three phase flow
Can. J. of Chem. Eng., 50, p.323-336

Boutin, P. and Wheeler, D.A. (1967) Column flotation, Mining World, 20(3), p. 47-

Castillo, D.I., Dobby, G.S. and Finch, J.A. (1988) Fine particle separation
performance in a flotation column under conditions of heavy froth loading, in
Production and Processing of Fine Particles (A.J. Plumpton, ed.), CIM Conference
of Metallurgists, Montreal, Pergamon, p.169-180

Cienski, T. and Coffin, V.L. (1981) Column flotation operation at Mines Gaspe
molybdenum circuit, Proc. 13th Annual Meeting of the Canadian Mineral Processors,
January, p.240-262

166
Clingan B.V. and McGregor, D.R. (1987) Column flotation experience at Magma
Copper Co., Minerals and Metallurgical Processing, 3(3), p.121-125
Coffin V L. and Miszczak, J. (1982) Column flotation at Mines Gaspe, Proc. 14th
Int. Mineral Processing Congr., Toronto, Canada, Paper 4-21
Contini N J Wilson, S.W. and Dobby, G.S. (1988) Measurement of rate data in
SoTon column! in Column Rotation '88 (K.V.S. Sastry, ed.), SME Annual
Meeting Phoenix, Arizona, p.81-89
Cova, D.R. (1974) Axial mixing in the liquid phase in gas-sparged columns, I&EC
Process Design and Devel., 13(3), p.292-296
Crawford, R. and Ralston, J. (1988) The influence of particle size and contact angle
in mineral flotation, Int. J. Mineral Processing, 23, p. 1-29
Deckwer, W.D., Burckhart, R., Zoll, G. (1974) Mixing and mass transfer in tall
bubble columns, Chem. Eng. Sci., 29, p.2177-2188
Dell, C.C., Bunyard, M.J., Rickleton, W.A. and Young, P.A. (1972) Release
analysis: acomparison of techniques, Trans. Inst. Min. Metall. (Section C), 81 (787)
Dell C C and Jenkins, B.W. (1976) The Leeds notation Column, Seventh Int.
Coal Preparation Congress, Sydney, Australia (A.C. Partridge, ed.), paper J3
del Villar, R. (1987) private communication
del Villar, R., Finch, J.A., Yianatos, J. and Laplante, A.R. (1988) Column flotation
simulation in Computer Applications in the Mineral industry, First Canadian
Conference, (K. Fytas, J. L. Collins and R. K. Singhal, eds.) A. A. Balkema,
Rotterdam
Dippenaar, A. (1982) The destabilization of froth by solids 1. The mechanism of
film rupture, Int. J. Mineral Processing, 9, p.1-14
Dirsus, D. F. (1988) Flotation columns: level sensing in three-phase slurries, B.Sc.
Thesis, Geo-Engineering, University of Toronto
Dobby, G.S., (1984) A fundamental flotation model and flotation column scale-up,
Ph.D. Thesis, McGill University, Montreal, Canada
Dobby, G.S. and Finch, J.A. (1985) Mixing characteristics of industrial flotation
columns, Chem. Engng. Science, 40(7), p. 1061-1068
Dobby, G.S., Amelunxen, R. and Finch, J.A. (1985) Column flotation: some plant
experience and model development, in Symp. Automation for Mineral Resource
Devel., Brisbane, Australia, p.259-264
Dobby, G.S. and Finch, J.A. (1986a) Flotation column scale-up and modelling,
C.I.M. Bulletin, 79 (889), p.89-96
Dobby, G.S. and Finch, J.A. (1986b) A model of particle sliding time for flotation
sized bubbles, J. Coll. Inter. Sci., 109 (2), p.493-498

167
Dobby G.S. and Finch, J.A. (.987) Panicle size dependences flotation derived
p 241 260 °f ^ CaP'Ure Pr0CeSS' 'nt-J- Mineral Processi"8- 2'
Dobby G.S., Yianatos, J.B. and Finch, J.A. (1988) Estimation of bubble diameter
T2 S fr°m dHft "UX analySiS"Canadian Metallur^icaI Q™££
Egan, J.R., Fairweather, M.J. and Meekel, W.A. (1988) Application of column
flotation to lead and zinc beneficiation at Cominco, in Column Flotation "88
(K.V.S. Sastry, ed.), SME Annual Meeting Phoenix, Arizona, p. 19-26
Englebrecht, J.A. and Woodburn, E.T. (1975) The effects of froth height, aeration
rate and gas precipitation on flotation, J. S. Afr. Inst. Min. Metall., 76. p. 125-132
Espinosa-Gomez, R. (1987) private communication
Espinosa-Gomez, R. (1989) Recovery of slimes discard at Niobec by column
flotation, Ph.D. thesis, McGill University, Montreal, Canada
Espinosa-Gomez, R., Finch, J.A. and Johnson, N.W. (1988a) Column flotation of
very fine particles. Minerals Engineering, 1(1), p. 3-18
Espinosa-Gomez, R., Finch, J.A., Yianatos, J.B. and Dobby, G.S. (1988b) Column
carrying capacity: particle size and density effects. Minerals Engineering, 1(1), p.

Espinosa-Gomez, R., Yianatos, J.B., Finch, J.A. and Johnson, N.W (1988c)
Carrying capacity limitations in flotation columns, in Column Flotation '88 (K V S
Sastry, ed.), SME Annual Meeting Phoenix, Arizona, p.143-148
Espinosa-Gomez, R.. Finch, J.A. and Bernert, W. (1988d) Coalescence and froth
collapse in the presence of fatty acid, J. Colloids and Surfaces, 32 (3/4), p. 197-210
Falutsu, M. and Dobby, G.S. (1989) Direct measurement of collection zone
recovery and froth drop back in a laboratory flotation column. Minerals Engineering
in press 6 6'

Feasby, D.G. (1987) editor. Column Rotation Cell Symposium Proceedings Trail
B.C., Canada, published by CANMET, Ottawa, June 1988 '
Feeley, CD., Landolt, C.A., Miszczak, J. and Steenburgh, W.M (1987) Column
flotation at INCO's matte separation plant, pres. 89th AGM-CIM, Toronto Canada
May

Finch, J.A., Yianatos, J.B. and Dobby, G.S. (1989) Column froths, in Frothing in
Flotauon: The Jan Leja Volume (J. Laskowski, ed.) Gordon and Breech, in press
Flint. L.R. and Howarth, W.J. (1971) The collision efficiency of small particles
with spherical air bubbles, Chem. Eng. Sci., 26, p. 1155-1168 ~
168
Bint IM MacPhail,P.andDobby,G.S.(1988)Aerosolfrotheraddn1onlnco.umn

flotaiion,C.I.M. Bulletin, 81(913), p. 81-84

and fluorite, Canadian Metallurgical Quarterly, 25 (1), p.15-21


Furey J. (1986) private communication
Harris, C.C. (1976) Flotation machines, in Botation-A.M. Gaudin Memona,
Volume (M.C. Fuerstenau, ed.) A.I.M.E., N.Y., p.753-815
HofferU.(1987) in Column notation Cell Symposium Proceedings (Feasby.ed.),
tS B.C., Canada, published by CANMET, Ottawa, June 1988
Huls B (1988) private communication

Jameson, G.J. (1986) private communication


Jameson G J. (1988) A new concept in flotation column design, in Column
Son '88 (K.V.S. Sastry, ed.), SME Annual Meeting Phoemx, Anzona, P.281-

Jiang, Z.W. and Ho.tham, P.N. (1986) f^^


particles and bubbles in flotation, Trans. I. M. M., 95, p.
Johnson, N.W. (1988) private communication
Jowett A (1980) Formation and disruption of particle-bubble
sruption of particlebubb aggregates-.m
ggg
rS^Vil
rSon^FineVarticles P~ Proc,*
Proc.,* Symp on Fme ParUC.es, Las
Vegas, Nev. (P. Somasundaran, ed.), AIME, N.Y., p.720-754
Klassen, V.I. and Mokrousov, V.A. (1963) An Introduction to the Theory of
Flotation, Butterworths, London, Chap. 5
Kava M and Laplante, A.R. (1988) Evaluation of the potential of wash water

October 4-6, p.175-186


Kosick G A Freberg, M. and Kuehn, L.A. (1988) Column flotation of galena at
the Polaris concentrator, C.I.M. Bulletin, 81 (920), p. 54-60
Laolante AR Toguri, J.M. and Smith, H.W.( 1983) The effect of air flow rate on

Int J. Mineral Processing, 11, p-203-219


Laplante, A.IL,Yta-lo..J3.«rfI»JA.(l«»)^^.^^^
ofLcollection zone in flotation columns, .n Column,Flotat.on 88 (K.V.S. Sastry,
ed.), SME Annual Meeting Phoenix, Arizona, p.69-80

169
Levenspiel. o. (1972)-Chemical Reacion Engineering", Wiley, N.V ChapIe,9

3SSS

170
columns Int. J. Mineral Processing, 23(3/4), p.265 m

., 51 (4),
efficiency, Canadian J. Chem.

486
Redfeam, M.A. (1986) private communication

Richardson, J.F. and Zaki, W.N. (1954) Sedimentation and fh.idizat.on: part 1,
Trans. Inst. Chem. Engrs., 32, p.35-52
Roscoe, R. (1952) The viscosity of suspensions of rigid spheres, Br. J. Appl. Phys.
3, p.267-269 .
Sastry, K.V.S. (1988) editor, proc. Column Flotation '88, SME Annual Meetmg
Phoenix, Arizona
G (1988) Design and operation of the Hydrochem

Meeting Phoenix, Arizona, p.287-292


H J a977)NewtheOreticalandexperimentalinvestigationsofstab.htyof
eticalandexp

SSrigUa.e
SriUa.es in flotation: a theory on *e upper Part.c,e s.ze of
floatability, Int. J. Mineral Processing, 4, p.241-259
' k G (1979) Investigations of the hydrodynamic

N.Y.p.63-84
Seeley, L.E., Hummel, R.L.

Stan, Y.T., Kelkar, B.G., Godbole, S.P.


W r>

ed,). Wished by ZSS&g. ^ %°°"^ - C. Ban*Ma,

l K.L. (1,48, Klne,ics „, Ae „„«„„ process.,. ^ Col|. &i K

ili. K.R. (,9,7) Co,umn «HS rai« ^ , Sieniu, E

Szatkowski, M. and Freybereer W r (wsi^ v ■


bubb,«s, Tran, ^ M^. £,^^^7^70

JWIk G.B. (19691 ..One Dimensiona,

^172
, DA (1983) Column flotation semmar, McG.ll
Wheeler
Canada, May column, in Froth Flotation,

7*

Yianatos, J.B.,
Processing. 15. p.279 .
flotation froths.
173''
^
i R A

- LT^^L^^ "98n8) U"""" P"*"™ and


Z2X
f~n^
andR.D. Caudle, eds.) sanautIllza"°n of High Sulfur Coals II (Y.P.Chugh
Yoon,R.H.,Adel,G.T.andLuttrell GH fHttftWK
process and apparanis for separa ing fiieS,^S-^entaPPlno-5761008, A

achine for meaHic. TOmSS^TT'*- "' "" Fl<"llire


^ .,, SME ZT^X^"2c^

174
INDEX
Bubble Reynolds number 46

Bubble rise velocity 16


Aerosol 34
Alignment, misalignment 65
Bubble size
Alternative column designs 1 carrying capacity 91-9

Angle of contact
carrying rate 25-26
definition 45 collection 50-51
maximum 44 entrapment 81"82
Amenability testing 119- estimation 17-18,20
Apparent gas holdup 1M mean 18
Attachment 38-^> .-^7
measurement 20
Attachment efficiency 38-39,43-4 profile 76

Mial dipersion coefficient 63-65


plug flow dispersion 60 Bubble surface._
Bubble swarm 18,--,-
Backmixing 65 Bubbly flow
Baffles 32-33,68.72.116 definition 13
model 17-26
Bank
recovery 104 theory 17-26
versus column 103-108
Base unit 112-113 CANMET 6
Carrying capacity
%£ 152-153, 155.157 definition 91-95
effect of column diameter %
definition 148-149
effect of variables 15J
effect of particle size 92-96
example calculation 115-116
cleaning 86-88
froth stability 75-76 measurement 92-96,126
gas holdup 82-84 model 91,93-96
Lformance 87,134-137 operating implications 97
measurement 149-151 ■ -rate 26-26,127
negative 80-81,100

positive 3
ratio 148 »25
Boundary conditions 61 Chum-turbulent 13-14.26
Bubble bed Cleaning 5,121.130-131,133
expanded 7o-"
packed 76-78 control 158-159
Bubble diameter (see Bubble size)
design 97,130-134
parallel 131-133
Bubble-loading review 130-134
scavenging 97,130-134
124-125
Bubble/particle attachment series 97,131-133
(see Attachment) simulation 127-134
Bubble/particle collision
Cleaning _ 175'
(see Collision)
effect of bias water 86-88
effect of bubble size 84-86 optimizing 88, 156-158
effect of froth depth 88 Process matrix 152-154
effect of gas rate 84-86 stabilizing 154-155
wash water 157

drop back 103


Copper 5-6, 87, 90, 120 122
'29-130, 156-157
froth 75-78,97-100 c"t«cal (particle) size 39
model 98-99
quantification 20-^> Definitions 4
Cocurrent Degree of mixing 60
low 16, 109 Density
operation 89, 109
bubble-panicle aggregate
Co Me (see Froth stability)
/fJ

Co'ecnon (see also Model) 37-57 froth zone 139, I41-143


Co iecnon efficiency 39,48-54
Co»ect.on rate constant 37.3584
Collection zone basic 1-3
flow regimes 13 case studies P7.n9
height, definition 68-69 circuit 97, 130-135"
ne.ght to diameter ratio 69-79- simulation II 1-1,5
-e-nonwi.thfh Detachment, disruption 39 89
Dm,en.on, seea,oNo^
mixing 59-73 radial distance 44-45
Particle attachment 43-48 time 40
Particle collection 37 57 velocity 44-45
ra.econ.s,ams 59-62, ,07-109/ Disjoining fijm 43

recovery 59-62. 103 111 M< DiSPe 6365

^'^ffiW Dispersion number


Cominco 5-6 13-5

C
95

Dowfroth 14, 21 81
Conductivity Drag 39
bias 150 Drift flux 18

D
level 146-147
measurement 146-147
Profiles 150
Contact angle 98-99 Efficiency
Control
attachment 47-48
bias 157 collection 39 '
circuits 158-159 collision 41-43
'eve! 88, 158
effect of solids 24-25
effect of sparger 27-32
Entrainment
effect of bias 86-88 froth zone 76.78-85
effect of bubble size 84-86
local 9
effect of froth depth 88 overall 9
effect of gas rate 84-86 predicted 17-19
profiled 75.78-85
Falconbridge 121-123, 129 measurement 9-13, 151
Gas rate .
location 68-69 carrying capac.ty 90-91
percentsolids99,l34 collection 54-58, 1-4
use in control 153-154 collection zone gas holdup
Fi,mthinning(seealsoRupture) 43.76
13-15.17-19
First-order kinetics 38.59 control 153. 156-157
Float, level measurement w froth zone gas holdup 79-81

maximum 22-25,79-81
Flotaire 1
Free discharge 83 measurement 151
optimum 54-58
^"addition to gas phase 84-86 performance 124
bubble size 84-86 profiles 79-80
Gibraltar Mines 5-6,32.157-159
dosage units 99
Froth(s) (see also Froth.zone) Gold 6. 133
conventional 75, /»
Harbour Lights Mining 99,132
draining 75,78
flow properties 96 Heiaht 68-69
loading 98-99
stability 97-100
structure 75-78
p
84,98-99.103
Froth depth (see Level) Hydrophilic 84
Froth zone
carrying capacity Inco 6.157
(see Carrying capacity) induction time 43 47-58
drop back 89-91,109,112-114,
Interaction, zones 103-110

Infraction with collection zone Interface


loss 79-81
103-UO second 25
recovery (see Drop back)
selectivity 89-90 Kidd Creek Mines 95,157
Kinetics, first-order 59-61
Gas holdup
apparent lM- Laboratory column 116-119
collection zone 9-15 Launder water 83,U7
effect of frother 14 81-82
effect of gas rate 13-14. TO 81
effect of liquid rate 13-1D
177
Les Mines Gaspe 5 89
Level Nickel 120-126, 129
control 88, 153-155 Niobec 21, 157
measurement 88, 138-147 Normalized
Liquid interstitial velocity 60 68 conductivity 146-147
L-quid vessel dispersion number lemperature 77. 144-146
(see Vessel dispersion number) Overall recovery 107-109
Magma Copper 156
Manganese dioxide 66-77 Parameter estimation
Manometers (see Pressure) carrying capacity 122,126
frOth 7Ano -A
Mass recovery 113-114
McGill University 5
Mechanical cell(s) rate constants 123-126
Panicle/bubble attachment
(see Attachment)
versus column 103-107 Panicle/bubble collision
"9-123,132 (see Collision)
MIBC 81
Mixing (see also Transport)
backmixing 65 carrying capacity 90-96
collection zone 59.73 collision, attachment 48-49
degree of 60 critical 39
effect of alignment 65-67 drop back 92
effect of baffles 67-68 distribution 94
"i froth 83 residence time 66-67
models 59-61 Panicle sliding
recovery 60-62, ,,3-1,4 model 44-47
time 44-47
attachment 43-48 Panicle slip velocity
carrying capacity 91,93, 96 (see Slip velocity)
carrying rate 25-26 Particle trajectory 42-44
coalescence 98-99 vessel dispersion number
Particle67-68
collection 38-48
collision 39-43 Peclet number 60
mixing 59-61 Pentlandite 120, 122
Panicle trajectory 40, 43-44 Pilot plant 116-119,~p6-127
Plug flow dispersion, axial 60-63 129-130 "
scale-up Ill-lie Plug flow
tanks-in-series 59, 105 dispersion coefficient 60-63 -
iranspon 59-61
m 1 Jf^imeraction '03-110 Polaris 6, 132, 157-159
Molybdenum 5-6,89-90
Mount (Mt.)Isa Mines 6 93-95 Potential flow 42-43,45.47 "'"
120-126, 157-159 Pressure
J78
levelmeasucynem 88, 138-143
Selectivity 89-90. 103-109
manometers 10-12
Simulation
measurement 10-13
circuit 116. 127-134
Process matrix 152-154
height to diameter ratio 69-73
Profiles
scale-up 111-116
electrical conductivity 150
zone interaction 103-105.
gas holdup 79.82, 150
112-114
temperature 76-77, 150
Sliding time 44.47
Pyrrhotite 125
Slip velocity
bubble 16-20.22
Rate constant
panicle 66-67. 113
collection 37-385, 105-109.
Solids
111-115
gas holdup 25
definition 107-109
froth stability 97-100
first-order 38
loading 98
measurement 123-126
Spargers
non first-order 109
internal 1.27-34
overall 107-109.111
external
Reactor
frother addition to 33-34
perfectly mixed 60
flexible 27-32
plug flow 60, 62
hole density 29-32
real 60,62
hole size 29-32
Recollection 103. 109
multiple 32-33
Recycle
rigid 27-32
between columns 132,134
sinale 29-32
between zones 103-106
surface area 29-32. 110
Reduced (see Normalized)
Stokes flow 41.43,45-46
Release analysis 3
Stokes number 40-43
Residence time
Strathcona (see Falconbridge)
liquid 64-65,68,112-114
Superficial velocity (or rate).
nominal 118, 125
definition 4
particle 38,65-68,104,
Surface area rate (see also
112-113,125
Bubble surface rate) 81
Residence time distribution
Surface vorticity 45-46
mean 61
measurement 61-63, 118 Symposium 6

variance 61
Tangential velocity 44-46
Retention time (see Residence time)
Tanks-in-series 59. 105
RTD (see Residence time distribution)
Temperature
Rupture 43,76,99
bias 150-151
level 144-146
Scale-up
profiles 150-151
case studies 120-130
Terms 4
example 112-115
TEB 21,81
model 111-116

179.
Tracers 60-63, 66, 84-85
Transport
perfectly mixed 59, 105-109
plug flow 59, 107-109
Turbulent mixing (see Mixing)

University of Toronto 95

Variance 61
Velocity
interstitial 60
slip (see Slip velocity)
tangential 44-46
Vessel dispersion number 60-68 71
113
Viscosity 25,52-53,111-113

Wash water
control 152, 155, 157
distributor design 83
distributor location 78, 83-84
mechanical cells 72
rate 86-87, 118
use of warm water 87
Water recovery 84-86

Zinc (see Lead)

You might also like