You are on page 1of 17

Int. J. Miner. Process. 64 Ž2002.

135–151
www.elsevier.comrlocaterijminpro

The modelling of froth zone recovery in batch and


continuously operated laboratory flotation cells
M.A. Vera a,) , Z.T. Mathe b, J.-P. Franzidis a , M.C. Harris b,
E.V. Manlapig a , C.T. O’Connor b
a
Julius Kruttschnitt Mineral Research Centre, The UniÕersity of Queensland, Brisbane, QLD 4068, Australia
b
Department of Chemical Engineering, UniÕersity of Cape Town, Cape Town, South Africa
Received 1 March 2001; received in revised form 15 June 2001; accepted 2 July 2001

Abstract

This paper proposes an integrated methodology for modelling froth zone performance in batch
and continuously operated laboratory flotation cells. The methodology is based on a semi-em-
pirical approach which relates the overall flotation rate constant to the froth depth ŽFD. in the
flotation cell; from this relationship, a froth zone recovery Ž R f . can be extracted. Froth zone
recovery, in turn, may be related to the froth retention time ŽFRT., defined as the ratio of froth
volume to the volumetric flow rate of concentrate from the cell. An expansion of this relationship
to account for particles recovered both by true flotation and entrainment provides a simple model
that may be used to predict the froth performance in continuous tests from the results of laboratory
batch experiments. Crown Copyright q 2002 Published by Elsevier Science B.V. All rights
reserved.

Keywords: froth flotation; froth zone recovery; froth depth; entrainment; water recovery

1. Introduction

Two distinct phases prevail in a flotation cell: a pulp phase and a froth phase. The
pulp phase has been studied extensively; however, it is only recently that the froth phase
has been recognised as a region which contributes significantly to the final flotation
performance.
It is in the froth phase that the essential processes of particle transport to the
concentrate launder and drainage to remove gangue entrainment occur. Many of the

)
Corresponding author.
E-mail address: m.vera@jktech.com.au ŽM.A. Vera..

0301-7516r02r$ - see front matter. Crown Copyright q 2002 Published by Elsevier Science B.V. All rights
reserved.
PII: S 0 3 0 1 - 7 5 1 6 Ž 0 1 . 0 0 0 6 8 - 0
136 M.A. Vera et al.r Int. J. Miner. Process. 64 (2002) 135–151

novel flotation technologies developed since the 1980s are aimed at improving methods
of contacting air bubbles and treated mineral particles in the pulp phase. However,
through the introduction of column cells Žoperated with deep froths, frequently with
water addition. and the design of booster plates and booster cones to assist with froth
transport in large tank cells, the scientific and engineering community has acknowledged
the importance of the froth phase in determining overall flotation performance.
This paper reviews recent work carried out at the Julius Kruttschnitt Mineral
Research Centre at the University of Queensland and the Mineral Processing Research
Unit at the University of Cape Town to measure and model froth zone recovery, as part
of the Australian Mineral Industries Research Association ŽAMIRA. P9 project. A
methodology for determining froth zone recovery is described; a model of froth zone
recovery based on froth retention time ŽFRT. is presented; and the model is applied to
batch and continuously operated laboratory flotation cells.

2. Froth zone performance

The metallurgical performance of the froth phase in a flotation cell may be expressed
in terms of froth zone recovery, which is frequently defined for particles recovered by
true flotation only Žto distinguish between the particles recovered by entrainment.. The
mass transfer between the pulp Žcollection. zone and the froth zone for particles that are
floated by true flotation is shown schematically in Fig. 1. R c is the collection zone
recovery due to true flotation Žthis does not include recovery due to entrainment.; R f is
the froth zone recovery, with respect to solids entering the froth attached to gas bubbles.

Fig. 1. Interaction between zones in a flotation cell.


M.A. Vera et al.r Int. J. Miner. Process. 64 (2002) 135–151 137

Mathematically the overall recovery R due to true flotation Žwhich does not include
recovery due to entrainment. can be described by the following equation ŽDobby, 1984;
Falutsu and Dobby, 1989.:
Rc Rf
Rs . Ž 1.
Rc Rf q Ž 1 y Rc .
The fraction of the solids entering the froth attached to gas bubbles that returns from
the froth to the pulp is known as the drop-back. The froth zone recovery Ž R f . is its
complement, which is defined as the recovery to the concentrate of particles entering the
froth zone from the collection zone attached to gas bubbles. R f may be calculated as the
total rate of transfer from the pulp to the concentrate divided by the rate of transfer from
the pulp to the froth phase:
k
Rf s . Ž 2.
kc
In many cases, the recovery of valuable mineral in a flotation feed by entrainment is
a small proportion of its recovery by true flotation. This is true, for example, for most
base mineral flotation plants, where the grind is not ultra-fine Žless than 10 mm.. In
these cases it is reasonable to calculate k, the overall first-order flotation rate constant,
from the measured overall recovery R by considering the system as a perfect mixer:
R
ks Ž 3.
t Ž1 yR.
where t is the mean residence time. R f may also defined independently of any mixing
regime as follows ŽSavassi, 1998.:
Mass rate of particles reporting to the concentrate via true flotation
Rf s . Ž 4.
Mass rate of attached particles at the pulp–froth interface
The effect of R f on the overall recovery R is clearly significant ŽEq. Ž1... A particle
may be collected in the pulp zone and report to the froth zone. However, if the recovery
in the froth zone R f is low then the particle would most likely not be recovered into the
concentrate but would drop back to the pulp. To illustrate the effect of R f on R, a
3D-plot of R as a function of R f and R c is presented in Fig. 2. This shows that the froth
zone recovery Ž R f . Žor its complement, drop-back. is a good measure of the loss of
recovery in the froth zone. This loss of recovery in the froth will diminish the overall
flotation recovery, R. Experience shows that collection zone recovery Ž R c . varies from
60% to 99%, while froth zone recovery Ž R f . varies from 10% to 90%. Consequently,
according to Fig. 2, up to 50% of the overall recovery Ž R . might be lost owing to
inefficiencies in the froth zone performance. This simple observation is indicative of
how detrimental a neglected froth phase can be to the overall flotation performance.
Unfortunately, until recently, the practical determination of R f has proved difficult.
At the pulp–froth interface, particle transport occurs in both directions, in a dynamic
fashion, complicating attempts at measurement. Additionally, much of the drop-back
from the froth to the pulp occurs in the first few centimeters of the froth zone, where
138 M.A. Vera et al.r Int. J. Miner. Process. 64 (2002) 135–151

Fig. 2. Overall recovery as function of froth zone recovery and collection zone recovery. 3D-plot of Eq. Ž1..

rapid deceleration of the bubble and initial bubble coalescence both occur. However, the
measurement and prediction of R f are of critical importance to the ultimate goal of
modelling and simulation of the flotation process.

3. A methodology for determining froth zone recovery


Three techniques for measuring R f in a flotation cell have been developed as part of
the AMIRA P9L project ŽAMIRA P9L Final Report, 2000.. Of these, the one that will
be described in this paper is the simultaneous determination of the collection zone rate
constant k c and the froth zone recovery, R f . This technique was derived by Vera et al.
Ž1999a,b. from previous work by Feteris et al. Ž1987. and Laplante et al. Ž1983..
The procedure is based on sample collection at different froth depths, keeping the
pulp zone volume constant. Two important assumptions are involved in this methodol-
ogy: Ž1. that the transfer of particles from the body of pulp to the pulprfroth interface
depends only on events occurring in the pulp zone, and Ž2. that the transfer of particles
from the froth to the concentrate depends only on events occurring in the froth zone. As
this methodology has been described previously by the authors ŽVera et al., 1999b., only
a brief resume is presented.
The method hinges upon the further development of Eq. Ž2. by incorporating a
relationship between flotation rate constant Ž k . and froth depth ŽFD.. The existence of a
linear relationship with negative slope between flotation rate constant and froth depth is
well known Že.g. Engelbrecht and Woodburn, 1975; Laplante, 1980; Feteris et al., 1987;
Hanumanth and Williams, 1990.. This relationship can be expressed as follows:
k s a y b Ž FD . Ž 5.
where a and b are constants. Now, when froth depth is zero, k c , the collection rate of
the pulp zone, may be determined by extrapolation as shown in Fig. 3.
M.A. Vera et al.r Int. J. Miner. Process. 64 (2002) 135–151 139

Fig. 3. Flotation zone rate constant Ž k . as a function of froth depth ŽFD..

Hence, a s k c . Next, if ŽFD. ks0 is the intercept of the straight line ŽEq. Ž5.. with the
X-axis, i.e. when k s 0 Žthis is equivalent to having a very deep froth, so that no
material is transferred from the froth to the concentrate., b s k crŽFD. ks0 . Substitution
for a and b into Eq. Ž5. leads to:
Ž FD.
k s kc 1 y½ Ž FD. ks 0 5 Ž 6.
which, when substituted into Eq. Ž2., yields:
Ž FD.
½
Rf s 1 y
Ž FD. ks 0
. 5 Ž 7.
The methodology for determining R f in a particular flotation cell consists in
operating the cell at a number of different froth depths, keeping the pulp volume
constant. Overall recovery of the mineral of interest is measured at each froth depth,
from which the flotation rate constant at each froth depth is obtained from Eq. Ž3..
Values of k are plotted against froth depth as in Fig. 3, and k c and ŽFD. ks0 are obtained
by extrapolation. This allows the calculation of R f at the experimental conditions, or
indeed any froth depth, using Eq. Ž7.. As part of the AMIRA P9 project, this procedure
has been carried out in 3- and 16-l laboratory cells, a 3-m3 pilot cell and industrial scale
flotation cells up to 100 m3 in volume ŽAMIRA P9L Final Report, 2000.. The effect of
flotation operating variables on R f may be investigated in this way, as described in the
section below.

4. Effect of operating variables on froth zone recovery


The above-described procedure was used to study the effects of operating variables,
such as air flow rate, impeller speed, percent solids, wash water rate and collector and
frother doses, on k c and R f of chalcopyrite and pyrite in the copper concentrator at
Mount Isa Mines ŽQueensland, Australia.. The work was carried out using the 16-l
JKMRC high S b flotation cell, details of which have been reported previously ŽVera et
140 M.A. Vera et al.r Int. J. Miner. Process. 64 (2002) 135–151

al., 1999a.. The most important feature of the JKMRC high S b cell, which distinguishes
it from almost all other laboratory scale sub-aeration flotation cells, is the air injection
system, which consists of a static in-line mixer of the type developed by Yoon et al.
Ž1989.. An aerator like this is capable of generating very fine bubbles which allow a
much higher bubble surface area flux Ž S b . to be produced, similar to those that have
been measured in full-scale flotation cells ŽVera et al., 1999a,b.. This enables the cell to

Fig. 4. Effect of operating variables on collection zone rate constant Ž k c . and froth zone recovery Ž R f . of
chalcopyrite ŽCPY. and pyrite ŽPY., at 5 cm froth depth. Ža. Collection zone rate constant and froth zone
recovery versus air flow rate Ž%solidss 20, impeller speeds1070 rpm, wash water rates 290 cm3 rmin,
collector doses 240 grt, and frother doses8 ppm.. Žb. k c and R f versus solids percent ŽAFR 52 lrmin,
ISs1070 rpm, WWFR s 290 cm3 rmin, collector doses 240 grt, and frother doses8 ppm.. Žc. k c and R f
versus impeller speed ŽAFR 52 lrmin, %Sols 20, WWFR s 290 cm3 rmin, collector doses 240 grt, and
frother doses8 ppm.. Žd. k c and R f versus wash water flow rate ŽAFR 52 lrmin, ISs1070 rpm, %Sols 20,
collector doses 240 grt, and frother doses8 ppm.. Že. k c and R f versus collector concentration ŽAFR 52
lrmin, ISs1070 rpm, WWFR s 290 cm3 rmin, %Sols 20, and frother doses8 ppm.. Žf. k c and R f versus
frother concentration ŽAFR 52 lrmin, ISs1070 rpm, WWFR s 290 cm3 rmin, collector doses 240 grt,
%Sols 20. ŽVera et al., 1999b..
M.A. Vera et al.r Int. J. Miner. Process. 64 (2002) 135–151 141

be used directly for scale-up Žwithout having to extrapolate rate constants to larger S b
values.. The cell is also equipped with a porous ceramic plate through which wash water
is added, to eliminate Žor at least minimise. recovery by entrainment; hence, the
calculated R f is for attached particles only, i.e. those recovered by true flotation.
The results of this study are summarised in Fig. 4. As expected, air flow rate had a
strong effect on the collection zone rate constant k c , proving once again the importance
of this parameter on flotation. The effects of percent solids, impeller speed and collector
dose on pulp zone performance were also important, i.e. these three variables affect the
mass transfer rate from the pulp zone to the froth zone. They appeared to have an
indirect influence on froth behaviour, to the extent that froth stability was altered. Froth
zone recovery R f was found to be strongly affected by impeller speed, air flow rate,
wash water addition, and frother dose Žand, of course, froth depth.. It is interesting to
note, however, that wash water addition and frother dose did not affect collection rate
constant. This supports the assumption that events occurring in the froth zone are not
influencing the pulp zone. Finally, in all the measurements conducted, froth zone
recovery was observed to be very similar in magnitude for both minerals species
analysed. This finding suggests that R f is non-selective for attached particles, i.e.
drop-back rates are equivalent for both minerals Žthis does not include entrainment..

5. Modelling of froth zone recovery


5.1. Particles recoÕered by true flotation
Floatable particles which arrive at the pulp–froth interface attached to air bubbles,
and remain attached, will report to the concentrate launder following the bubble flow
patterns. Froth zone recovery is not only a function of froth depth, as discussed above,
but also of froth area or froth transportation distance, which is a function of cell size and
design Žsee Fig. 5.. This is of special importance when scaling up metallurgical
performance from small to large flotation cells, and is one of the main reasons why it
has proved so difficult to scale up from laboratory scale to industrial scale cells.
More generally, R f will depend on the residence time of particles in the froth, which
is difficult to measure but may be approximated by the froth retention time ŽFRT.,
defined as the ratio of the froth volume to the concentrate volumetric flow rate. A
potentially very useful relationship between froth recovery of floatable minerals and
froth retention time has been observed in several flotation systems ŽGorain et al., 1998..
This is shown in Fig. 6, where it can be seen that the relationship between froth recovery
and froth retention time was essentially independent of cell size Ž0.25 and 2.8 m3 ., cell
mechanism ŽDorr-Oliver, Batequip and Outokumpu. or cell operating conditions Žfroth
depth, air flow rate and impeller speed..
On the basis of the relationship observed in Fig. 6 it is reasonable to suppose that the
froth zone recovery of particles which arrive at the pulp–froth interface attached to air
bubbles, and remain attached, f Ž a., can be described by a simple exponential function:
f Ž a . s exp Ž yb FRT . Ž 8.
where b is a parameter which is arguably related to the rate at which the bubbles are
coalescing and breaking-up, causing particles to detach and drain back through the froth
142 M.A. Vera et al.r Int. J. Miner. Process. 64 (2002) 135–151

Fig. 5. Schematic drawing of effect of froth depth and froth cross-sectional area on froth zone recovery Ž R f ..

towards the pulp. The detached particles, viz. Ž1 y f Ž a.., will either drain back into the
pulp or be recovered into the concentrate launder in the water between the bubbles, i.e.
via an entrainment mechanism. This mechanism would be expected to be particle

Fig. 6. Effect of froth retention time on froth recovery ŽGorain et al., 1998..
M.A. Vera et al.r Int. J. Miner. Process. 64 (2002) 135–151 143

size-dependent, and has been found to be the case ŽSavassi, 1998.. Fig. 7 shows the
observed effect of particle size on the R f –FRT relationship:
If the drainage rate of detached particles of size class i to the pulp phase, v i , is
proportional to the mass of these particles in the froth, the fraction remaining in the froth
after time t is expŽyv i t . ŽMoys, 1989.. Assuming that the detached particles are well
distributed throughout the froth phase Žmost likely in shallow froths., and using the
convolution integral approach, it can be shown that the fraction of detached particles that
would be entrained is ŽMathe et al., 2000a,b; Mathe, 2000.:
1
f Ž entri . s . Ž 9.
1 q v i FRT
This is similar to a classification function proposed by Bisshop and White Ž1976. to
describe entrainment of hydrophilic particles in bench-scale flotation cells:
1 q a FRT
CFi s Ž 10 .
1 q l i FRT
where a and l i are empirically derived parameters. In practice, the numerator in Eq.
Ž10. has been found to be very close to unity.
Combining Eqs. Ž8. and Ž9., the overall froth zone recovery of particles entering the
froth attached to bubbles Ži.e. by true flotation. can be expressed as:
1
R f Ž i. s exp Ž yb FRT . q 1 y exp Ž yb FRT . Ž 11 .
1 q v i FRT
where froth zone recovery R fŽ i. is now expressed on a size-by-size basis.

Fig. 7. Effect of particle size on the froth recovery–froth retention time relationship.
144 M.A. Vera et al.r Int. J. Miner. Process. 64 (2002) 135–151

5.2. Particles recoÕered by entrainment

It is well known that in parallel with the mechanism of true flotation, particles can
enter the base of a flotation froth entrained in the water between bubbles. Some of these
particles drain back into the pulp phase while others are recovered in the concentrate.
Unlike true flotation, which is chemically selective, both gangue and valuable minerals
alike can be recovered in this way. The degree of entrainment ENTi is defined as the
ratio of the entrainment recovery of solids within the froth Ži.e. the froth recovery of
particles by entrainment. and the recovery of water. This ratio is generally independent
of the water recovery but a strong function of particle size Žhence the subscript i
denoting size class..
It is reasonable to suppose that entrained particles will behave in the same way in the
froth phase as particles that become detached from bubbles, i.e. their behaviour will also
be described by Eq. Ž9.. Hence, the froth zone recovery of particles entering the froth
suspended in the water between bubbles Ži.e. by entrainment. can be expressed as:
1
s ENTi R w Ž 12 .
1 q v i FRT
where R w is the recovery of water from the cell. The practical use of Eqs. Ž11. and Ž12.
in the analysis of batch and continuous froth performance is now considered.

6. Prediction of continuous results from batch data

The laboratory batch test can be considered the workhorse of flotation research and
development. It is used in fundamental flotation studies, for reagent development and
evaluation, for everyday plant performance testing, and plant optimisation as well as for
flotation plant design. Correlation between batch and continuous tests is known to be
poor, for several reasons, the most important of which is the transportation of particles
in the froth phase: in continuous tests, the froth characteristics are at steady state, while
in batch tests the character of the froth changes constantly, as succeeding concentrates
are removed from the cell. However, through the use of a froth recovery model such as
the one proposed in Eq. Ž11. above, it should be possible to improve the correlation of
flotation performance between batch and continuous flotation systems substantially,
providing a basis for the prediction of the performance of one system based on results
obtained in the other.
In order to test this, a study was carried out using a modified Leeds laboratory-scale
flotation cell, operated in both batch and continuous modes. The cell had a nominal
volume of 3.5 l, a cross-sectional area of 207 cm2 and was fitted with a bottom-driven
impeller. High-purity quartz Ž) 99.6% SiO 2 . was used in the flotation experiments to
eliminate concerns about liberation and other problems associated with mixed mineral
ores. The material was milled and screened to 80% passing 106 mm then calcined at 500
8C for 2 h to drive off any impurities and organic materials coated on the surface of the
particles.
M.A. Vera et al.r Int. J. Miner. Process. 64 (2002) 135–151 145

Flotation experiments were carried out at 10% solids at a constant impeller speed of
1000 rpm. Operating conditions that were varied, so as to produce a range of metallurgi-
cal responses that would allow the modelling of both pulp and froth zone behaviour,
were froth depth and air flow rate, which were found previously to be the most
important variables affecting flotation response. Froth depth was varied from 0.5 to 5.0
cm, and air flow rate from 4 to 7 lrmin.
In the batch tests, samples were collected at cumulative flotation times of 30 s, up to
5 min. Froth removal was achieved by hand scraping at a constant rate of 6 scrapesrmin.
In the continuous flotation tests slurry was fed from a large conditioning tank fitted with
recirculation flow and a mixer. All other aspects of the operational procedure were
identical for both steady state Žcontinuous. and non-steady state Žbatch. tests. The
particle size distribution of all feed and concentrate samples was determined using a
Malvern particle size analyser.
An indication of the effect of the froth depth and the air flow rate on the flotation
response may be seen in Fig. 8, in which the cumulative recovery in the batch cell is
plotted against time for two pairs of experiments, at 0.5 and 4.5 cm froth depth, and 4
and 6 lrmin air flow rate. Overall recovery was much greater for the pair at the
shallower froth depth, with an increase in air flow rate resulting in an increase in
recovery from the start to the end of the experiment. At the deeper froth depth, recovery
was significantly less Ž50% less after 5 min at the lower air flow rate., but increasing the
air flow rate had a much greater effect on the performance. These results are typical of
those obtained in the study.

Fig. 8. Effect of the froth zone on the performance of the batch cell.
146 M.A. Vera et al.r Int. J. Miner. Process. 64 (2002) 135–151

In order to predict the performance of the continuous cell, a flotation model


incorporating the froth recovery model proposed in Eq. Ž11. above was fitted to the
batch test data. For this purpose, each batch test, which provides recovery-time
information under non-steady state conditions, was considered as comprising a sequence
of tests conducted under continuous, steady state conditions, with the residence time in
each stage corresponding to a concentrate collection period in the batch test. The overall
recovery of material in each period in a batch test was then considered to be a function
of the collection zone recovery, the water recovery, and the froth zone recovery of
particles recovered by true flotation and entrainment.
The overall flotation model fitted to the results was that developed by Savassi Ž1998.,
which incorporates recovery by true flotation and entrainment:
R cŽ i. R f Ž i. Ž 1 y R w . q ENTŽ i. R w Ž 1 y R cŽ i. .
R oŽ i. s Ž 13 .
Ž 1 y R w . Ž 1 y R cŽ i. q R cŽ i. R f Ž i. . q ENTŽ i. R w Ž 1 y R cŽ i. .
where R oŽ i. is the overall recovery of particles in size class i, R cŽ i. is the collection zone
recovery of particles in size class i due to true flotation, R fŽ i. is the froth zone recovery
of particles in size class i entering the froth attached to gas bubbles, R w is the water
recovery and ENTi is the degree of entrainment of particles of size class i as defined in
Eqs. Ž9. and Ž12.. It is readily apparent that if R w s 0 Žwhich effectively reduces
entrainment to zero. Eq. Ž13. collapses into Eq. Ž1., which is for particles recovered by
true flotation only.
In a batch flotation cell, R cŽ i. may be expressed as:
R cŽ i. s 1 y exp Ž yk cŽ i. t . Ž 14 .
where k cŽ i. is the collection zone rate constant as defined in Eq. Ž2. above, for particles
of size class i, and t is the flotation time. The work of Gorain et al. Ž1997. has shown
that the collection zone rate constant is a linear function of the ore floatability Pi and the
bubble surface area flux S b , i.e.:
k cŽ i. s Pi S b Ž 15 .
where S b is calculated from the equation:
6 Jg
Sb s Ž 16 .
db

where Jg is the superficial gas velocity and d b is the Sauter mean bubble size.
In the analysis of the batch data in this study, the ore is considered to be made up of
six floatability classes, corresponding to six particle size classes, with Žgeometric. mean
size 4, 17, 34, 57, 90, and 164 mm, respectively. The mass fraction of each of these
classes in the flotation feed and concentrates is known by measurement; the mass
fractions change from one flotation stage Žcollection time interval. to another, with the
tailings from each stage of the flotation test providing the feed to the next. The
floatability Pi ŽEq. Ž15.. of each particle size class is assumed to remain constant for the
entire batch flotation test, while the froth recovery factor, R fŽ i. , is assumed to be
M.A. Vera et al.r Int. J. Miner. Process. 64 (2002) 135–151 147

Table 1
Derived parameters from batch data
Size fraction Floatability, v Ž1rmin. b Ž1rmin.
Žmm. P Ž=10 3 .
Sub 10 1.3 7.7 4.0
10 to 25 1.2 6.4
25 to 45 1.6 10.2
45 to 75 2.0 19.7
75 to 106 2.3 35.5
plus 106 2.7 75.0

different in each stage of the batch flotation test, which in turn enables the extraction of
froth parameters that can be used to predict the froth performance in continuous tests.
The batch data were modelled using Eqs. Ž11. – Ž15. to provide estimates of the model
parameters Pi , v i and b , based on sum-of-error squares minimisation. Table 1 presents
the parameters derived. The model fit with respect to the batch data using the derived
parameters is shown in Fig. 9, on a size-by-size basis. It can be seen that the model can
describe quite adequately the performance of the batch cell over a wide range of
operating conditions and particle sizes. The model contains 13 parameters Žone value of
b , and six values Žone per size class. of Pi and v i . fitted to 250 data points.
Returning to Table 1, for the particle size range used in this study, both floatability Pi
and particle drainage rate v i increase with size, as expected. Closer examination reveals

Fig. 9. The fit of model parameters to the batch data.


148 M.A. Vera et al.r Int. J. Miner. Process. 64 (2002) 135–151

that the relationship between v i and mean particle size, d i , is well described by the
equation:

v i s a q bd i2 . Ž 17 .
In this equation, the parameter a can be considered to represent the water drainage
rate under the froth conditions arising in the study Žfrother type and concentration, solids
characteristics., which may be compared to b in Eq. Ž8., while parameter b in Eq. Ž17.
represents a proportionality constant for particle drainage rate as a function of size under
these froth conditions. If it is assumed that particles are settling predominantly via
bubble plateau borders, i.e. the intersections of the three bubble films, in which water is
draining over a distance h at a uniform rate Ž b . and the particles settle freely over this
distance, the b parameter can be assumed to be adequately described by the following
expression ŽWills, 1992.:

g Ž Ds y Df . d i2
bs Ž 18 .
18h h

where g is the gravitational acceleration, Ds is the density of the solid, Df is the density
of the fluid and h is the fluid Žfroth. viscosity, which is the only unknown that must be
fitted. After substituting Eqs. Ž15. and Ž17. into Eq. Ž11., the batch model was refitted to
the batch data. The derived a and b parameters were found to be 6.9 Žlrmin. and

Fig. 10. The revised model fit to the batch data.


M.A. Vera et al.r Int. J. Miner. Process. 64 (2002) 135–151 149

3.04 = 10y3 Žlrmin-mm2 ., respectively. An average value of the b parameter calculated


using the properties of water was found to be 3.20 = 10y3 Žlrmin-mm2 ., which was
very close to the fitted parameter.
Fig. 10 shows the revised model fit with respect to the batch data using the derived
parameters, on a size-by-size basis. Once again the model fits the performance of the
batch cell over a wide range of operating conditions and particle sizes. This time
however the model contains only eight parameters Žone value of b , one value of h , and
six values Žone per size class. of Pi . fitted to the same 250 data points. The sacrifice of
five parameters has not affected the model fit adversely.
The revised model parameters were used to predict flotation performance in the same
cell, using the same ore, under continuous, steady state operating conditions for a range
of different froth heights and gas rates. Fig. 11 compares the experimental and predicted
concentrate masses, on a sized basis, for these tests. It can be seen that the prediction of
the continuous cell performance using the model parameters derived from the batch test
data is very good. This is despite the fact that the froth stability parameter, b , derived
from the batch tests represents an AaveragedB characteristic froth drainage behaviour
over the duration of the batch test, given the fact that the stability of the froth decreases
with time, owing to the depletion of both frother and solids. Despite this apparent
limitation, the good agreement between predicted and measured concentrate masses by
size demonstrates the potential of this approach for relating the performance of batch
flotation data to flotation cells operated under continuous conditions.

Fig. 11. Prediction of continuous performance using batch derived parameters.


150 M.A. Vera et al.r Int. J. Miner. Process. 64 (2002) 135–151

7. Conclusions

This paper has highlighted the importance of froth zone in flotation, and has
described a technique for measuring froth performance in terms of froth zone recovery
factor, R f . Experiments using a 16-l high S b flotation cell indicated that R f was a strong
function of impeller speed, air flow rate, wash water rate and frother dose. A froth zone
recovery model based on froth retention time, defined as the ratio of the froth volume to
the concentrate volumetric flow rate, was fitted to laboratory batch flotation data and
used successfully to predict the performance of the same cell when operated continu-
ously.
At present, the primary potential application of this model is the prediction of the
froth zone performance of large industrial cells in plant operation, based on the results of
batch tests, as the only data required from the industrial cells are easily measurable
Žvolumetric flow rate of concentrate, pulp density of concentrate, froth volume, gas rate,
bubble size distribution.. Use of this method for design prediction is more limited, as a
predictive water recovery model for flotation is not yet available to allow the prediction
of FRT. However, for design prediction, the gas residence time Ž HrJg . can be
substituted for FRT, to provide a reasonable estimate of froth performance using the
proposed methodology.
The model as presented clearly offers considerable scope for refinement and further
research: for example, there is scope for evaluating the time dependence of the froth
stability parameter, b, in batch tests, and for determining an expression for b based on
measurable system properties.

Acknowledgements

The authors would like to thank AMIRA P9 project and its sponsors for the funding
which made this work possible.

References

AMIRA P9L Final Report, 2000. The Optimisation of Mineral Processes by Modelling and Simulation. Final
Report Number, Brisbane-Australia, Julius Kruttschnitt Mineral Research Centre, The University of
Queensland.
Bisshop, J.P., White, M.E., 1976. Study of particle entrainment in flotation froths. Transactions of the
Institution of Mining and Metallurgy Sect. C: Mineral Process. Extr. Metall. 85, C191–C194.
Dobby, G.S., 1984. A Fundamental Flotation Model and Flotation Column Scale-Up. Department of Mining
and Metallurgical Engineering. PhD thesis, McGill University, Montreal, p. 259.
Engelbrecht, J.A., Woodburn, E.T., 1975. The effect of froth height, aeration rate and gas precipitation on
flotation. Journal of The South African Institute of Mining and Metallurgy 76, 125–132.
Falutsu, M., Dobby, G.S., 1989. Direct measurement of froth drop back and collection zone recovery in a
laboratory flotation column. Minerals Engineering 2 Ž3., 377–386.
Feteris, S.M., Frew, J.A., Jowett, A., 1987. Modelling the effect of froth depth in flotation. International
Journal of Mineral Processing 20, 121–135.
M.A. Vera et al.r Int. J. Miner. Process. 64 (2002) 135–151 151

Gorain, B.K., Franzidis, J.P., Manlapig, E.V., 1997. Studies on impeller type, impeller speed and air flow rate
in an industrial scale flotation cell. Part 4: Effect Of Bubble Surface Area Flux On Flotation Performance.
Minerals Engineering 10 Ž4., 367–379.
Gorain, B.K., Harris, M.C., Franzidis, J.-P., Manlapig, E.V., 1998. The effect of froth residence time on the
kinetics of flotation. Minerals Engineering 11 Ž7., 627–638.
Hanumanth, G.S., Williams, D.J.A., 1990. An experimental study of the effects of froth height on flotation of
China clay. Powder Technology 60, 131–144.
Laplante, A.R., 1980. The effect of air flow rate on the kinetics of flotation. Department of Metallurgy and
Materials Science. Degree of Doctor of Philosophy thesis, University of Toronto, Toronto, Canada, p. 357.
Laplante, A.R., Toguri, J.M., Smith, H.W., 1983. The effect of air flow on the kinetics of flotation: Part 1. The
transfer of material from the slurry to the froth. International Journal of Mineral Processing 11, 203–219.
Mathe, T.Z., 2000. Modelling the influence of the froth phase on recovery in batch and continuous flotation
cells. Department of Chemical Engineering. PhD thesis, University of Cape Town, Cape Town, p. 220.
Mathe, T.Z., Harris, M.C., O’Connor, C.T., 2000a. Modelling the influence of the froth phase on recovery in
batch and continuous flotation cells. Proceedings of the XXI International Mineral Processing Congress,
23–27 July 2000. Elsevier, Rome-Italy, pp. B8-a33–B8-a39.
Mathe, T.Z., Harris, M.C., O’Connor, C.T., 2000b. A review of methods to model the froth phase in
non-steady state flotation systems. Minerals Engineering 13 Ž2., 127–140.
Moys, M.H., 1989. Mass transport in flotation froths. Mineral Processing and Extractive Metallurgy Review 5
Ž1–4., 203–228.
Savassi, O.N., 1998. Direct estimation of the degree of entrainment and froth recovery of attached particles in
industrial flotation cells. Department of Mining, Minerals and Materials Engineering, JKMRC. PhD thesis,
University of Queensland, Brisbane, p. 393.
Vera, M.A., Franzidis, J.P., Manlapig, E.V., 1999a. The JKMRC high bubble surface area flux flotation cell.
Minerals Engineering 12 Ž5., 477–484.
Vera, M.A., Franzidis, J.-P., Manlapig, E.V., 1999b. Simultaneous determination of collection zone rate
constant and froth zone recovery in a mechanical flotation environment. Minerals Engineering 12 Ž10.,
1163–1176.
Wills, B.A., 1992. Mineral Processing Technology, 5th edn. Pergamon Press, p. 855.
Yoon, R.H., Luttrell, G.T.A., Mankosa, M.J., 1989. Recent advances in fine coal flotation. In: Chander, S.,
Klimpel, R.R. ŽEds.., Advances in Coal and Mineral Processing Using Flotation. Society for Mining,
Metallurgy, and Exploration, Inc., Palm Coast, Florida, USA, pp. 211–218, December 3–8, 1989.

You might also like