You are on page 1of 12

Minerals Engineering 19 (2006) 871–882

This article is also available online at:


www.elsevier.com/locate/mineng

An evaluation of different models of water recovery in flotation


a,*
X. Zheng , J.P. Franzidis b, N.W. Johnson c

a
Newmont Golden Grove Operations, Department of Metallurgy, PMB 7, Geraldton, Western Australia 6530, Australia
b
Julius Kruttschnitt Mineral Research Centre, The University of Queensland, Brisbane, Australia
c
Mineralurgy Pty Ltd., Brisbane, Australia

Received 9 March 2005; accepted 28 July 2005


Available online 9 March 2006

Abstract

Water recovery is one of the key parameters in flotation modelling for the purposes of plant design and process control, as it
determines the circulating flow and residence time in the individual process units in the plant and has a significant effect on entrain-
ment and froth recovery. This paper reviews some of the water recovery models available in the literature, including both empirical
and fundamental models. The selected models are tested using the data obtained from the experimental work conducted in an
Outokumpu 3 m3 tank cell at the Xstrata Mt Isa copper concentrator. It is found that all the models fit the experimental data rea-
sonably well for a given flotation system. However, the empirical models are either unable to distinguish the effect of different cell
operating conditions or required to determine the empirical model parameters to be derived in an existing flotation system. The
model developed by [Neethling, S.J., Lee, H.T., Cilliers, J.J., 2003, Simple relationships for predicting the recovery of liquid from
flowing foams and froths. Minerals Engineering 16, 1123–1130] is based on fundamental understanding of the froth structure and
transfer of the water in the froth. It describes the water recovery as a function of the cell operating conditions and the froth prop-
erties which can all be determined on-line. Hence, the fundamental model can be used for process control purposes in practice. By
incorporating additional models to relate the air recovery and surface bubble size directly to the cell operating conditions, the fun-
damental model can also be used for prediction purposes.
 2005 Elsevier Ltd. All rights reserved.

Keywords: Froth flotation; Flotation bubbles

1. Introduction researchers and plant operators. This paper reviews


some of the water recovery models available in the liter-
Water recovery is one of the key parameters in flota- ature. The selected models are evaluated using data ob-
tion plant design and operation. It determines to a large tained from the test work conducted in an Outokumpu
extent the circulating load and residence time in the indi- 3 m3 tank cell in the Xstrata Mt Isa copper con-
vidual process units in the plant. Meanwhile, entrain- centrator.
ment and froth recovery are strongly associated with
water recovery and influence both mineral recovery
and concentrate grade. Hence, the ability to model water 2. Definition of water recovery
recovery and to control it, especially in industrial scale
flotation cells has always been an important goal of both Water recovery may be defined as the fraction of the
water entering the flotation cell that is recovered in the
*
Corresponding author. Tel.: +61 8 9956 4398; fax: +61 8 9956
concentrate. However, this definition of water recovery
4319. is only one of several found in the literature owing to
E-mail address: xiaofeng.zheng@newmont.com (X. Zheng). the use of different flotation systems by different workers

0892-6875/$ - see front matter  2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.mineng.2005.07.021
872 X. Zheng et al. / Minerals Engineering 19 (2006) 871–882

and use of the water recovery data for different pur- attempt was also made to model the water recovery
poses. For example, water recovery is sometimes calcu- based on fundamental understanding of the physics
lated as a fraction of the total water in the flotation cell (Moys, 1984). In the recent years, a significant progress
that is recovered to the concentrate. This definition of has been made with the fundamental approach (Neeth-
water recovery is commonly used for batch flotation ling et al., 2003). Some of the models for water recovery
tests. are now reviewed and evaluated with experimental data
It is also not unusual to use the water flow rate of the obtained from the test work conducted in an Outok-
concentrate directly instead of the fractional recovery of umpu 3 m3 tank cell in the Xstrata Mt Isa copper
the feed water or the total water in the cell, especially in concentrator.
a batch flotation cell when the volume of the pulp varies
during the test. Some researchers avoid use of the frac- 3.1. Water recovery as a function of solids recovery
tional recovery of water because the concentrate water
flow rate is believed to be relatively independent of the In this approach, concentrate water flow rate is linked
feed water flow rate or the amount of water in the pulp directly with the concentrate solids flow rate. This allows
phase. If so, it would be misleading to compare the frac- the water flows throughout the plant to be fixed un-
tional water recovery for different operating conditions iquely in flotation modelling (King, 1973). The model
or different sizes of flotation cells. In this instance, the does not require any information about the cell operat-
concentrate water flow rate would be a more proper ing conditions.
term to use. Based on the conservation of the volume and mass of
Finally, it should be noted that water recovery in a water and solids in a given stream, the relationship be-
flotation cell is the result of a two-step transfer pro- tween the water and the solids flow rates can be ex-
cess—transfer of the water in the pulp phase to the froth pressed as
phase and transfer of the water in the froth phase to the
SGs  SGp 1
concentrate launder. The amount of water entering the Qw ¼   Fs ð1Þ
froth phase from the pulp phase is closely related to SGp  SGw SGs
the pulp conditions such as the bubble surface area flux, or
the concentration of chemical reagents, the concentra-
1X 1
tion of suspended solids and the bubble loading condi- Qw ¼   Fs ð2Þ
tion. Only a certain fraction of the water entering the X SGw
froth phase may be recovered into the concentrate. where Qw, volumetric flow rate of the water; Fs, mass
The rest returns to the pulp phase via drainage and/or flow rate of the solids; X, mass fraction of the solids in
with the collapsed froth. These two steps of transfer the stream; SGs, SGp and SGw, specific gravity of the
are not totally independent of each other. The amount solids, slurry and water, respectively.
of water entering the froth phase may to a large degree Generally, the solids and water specific gravities can
determine the characteristics of the froth (froth mobility be assumed to be constant. Hence, to predict the water
and stability and solids/liquid/air hold-up in the froth flow from the solids flow is to predict the SG value of
phase) and hence influence the recovery of the water the concentrate slurry or the mass fraction of the solids
and solid particles in the froth phase to the concentrate in the stream. In practice, for control purposes, these
launder. values can be measured. The usefulness of the above
equations in providing a means of accurate control over
the process relies on how sensitive the concentrate slurry
3. Modelling of water recovery SG and solids content are to the process variables which
are used to control the process.
Research has been going for many years, trying to In JKSimFloat (Alford, 1990), a commercial com-
understand the froth behaviour and the mechanisms be- puter software package for flotation modelling and sim-
hind the water recovery in flotation. However, due to ulation, the concentrate water flow rate is expressed as a
the complicity of the froth system and the limitations power function of the concentrate solids flow rate:
in measurement techniques, water recovery was mod-
Qw ¼ a  F bs ð3Þ
elled simply by related to an certain aspect of the cell
performance, commonly used the solids recovery (King, where a and b are empirically fitted parameters.
1973; Alford, 1990), in the early stage. An alternative Again, the usefulness of this model depends on the
approach was to find some statistically significant corre- extent to which the model parameters, a and b, remain
lation between the water recovery and the cell operating constant under different cell operating conditions or in
conditions. Such an approach was widely adopted as it different cells.
provided a relatively quick and simple solution (Savassi, Table 1 summarises the concentrate water and solids
1998; Gorain et al., 1998; Harris, 2000). Meanwhile, flow rates measured in an Outokumpu 3 m3 tank cell in
X. Zheng et al. / Minerals Engineering 19 (2006) 871–882 873

Table 1 flow rates at the lower air rates. If the model is used to fit
Calculated concentrate water flow as a function of solids flow the data separately according to the air rate, the fitted
Air rate Froth %Solids Fs Qw Qw Qw model parameters for the different cell operating condi-
(l/min) height (%) (ton/h) (m3/h) (cal-1) (cal-2) tions now become:
(cm) (m3/h) (m3/h)
1085 9.6 36.66 1.69 2.92 3.37 2.95 a ¼ 0:94 and b ¼ 2:18 at the lower air rates
1085 14.3 45.53 1.17 1.40 2.13 1.32 a ¼ 2:09 and b ¼ 1:13 at the higher air rates
1085 19.0 48.27 0.92 0.99 1.58 0.78
1085 23.2 50.97 1.05 1.01 1.86 1.04 Fig. 1(b) shows that the fitting can be improved by treat-
1380 9.1 28.42 2.20 5.54 4.69 5.24 ing the data separately according to the air rate—the
1380 13.8 32.50 2.08 4.32 4.37 4.64
correlation coefficient improving from 0.986 to 0.993
1380 18.5 37.33 1.62 2.72 3.20 2.69
1380 23.2 39.12 1.60 2.49 3.15 2.62 and the residual sum square errors of the regression
1674 10.5 26.13 4.55 12.86 11.59 11.59 reducing from 6.154 to 3.205. However, it reveals the
1674 14.7 28.95 2.97 7.29 6.81 7.16 limitation of applying the model to a wide range of oper-
1674 19.4 28.90 2.61 6.42 5.80 6.19 ating conditions.
1674 24.1 32.30 1.77 3.71 3.57 3.99
Fig. 2 shows the concentrate water flow rates in each
1968 10.5 28.43 5.83 14.68 15.79 15.34
1968 14.7 29.26 3.77 9.11 9.17 9.37 of the three flotation cells of a rougher bank (Sala 80 l
1968 19.4 28.80 2.78 6.87 6.27 6.64 cells) in a platinum pilot plant (Harris, 2000). The data
1968 24.1 33.09 2.49 5.04 5.47 5.87 from the three cells were all fitted to Eq. (3) (Fig. 2(a))
Note: (cal-1)—modelling fitting using all the data; (cal-2)—model fit- and then fitted separately, cell by cell (Fig. 2(b)).
ting separately for different air rates. Fig. 2(a) clearly shows that the model fitted concen-
trate water flow rates in the first rougher cell fall in a dif-
ferent line from those in the second and third cells. The
the Xstrata Mt Isa copper concentrator. The experimen-
fitted constants a and b for the different cells are found
tal cell was fed with the plant rougher feed ore and oper-
to be:
ated at four different air rates and, at each air rate, four
different froth heights. Also included in Table 1 are the a ¼ 0:266 and b ¼ 1:08 for rougher cell 1
fitted results using Eq. (3). a ¼ 0:367 and b ¼ 1:05 for rougher cells 2 and 3
It can be clearly seen in Table 1 that the solids con-
centration in the concentrate varied significantly with The significant difference in the model constant ‘‘a’’ sug-
the cell operating conditions. Hence, the assumption of gests that the model derived from one flotation cell
a constant concentrate %Solids value in order to use Eq. might not be able to be used for other cells in the same
(2) for prediction purposes is not valid in practice. The circuit. In practice, feed at the beginning of the flotation
%Solids value or slurry SG has to be measured if it is bank contains a significantly higher content of floatable
used for control purposes. particles and a higher concentration of frother, which
If all the experimental data are used in the model fit- likely results in different froth properties from the cells
ting regardless of cell operating conditions (cal-1), the further down the bank. Therefore, it can also be ex-
fitted values of the constants a and b are: pected that the difference will remain in the cells of dif-
ferent circuits.
a ¼ 1:75 and b ¼ 1:25
Alford (1990) applied this model in flotation col-
It can be seen in Fig. 1(a) that there is a significant dis- umns. Uribe et al. (1999) studied the limitations of
crepancy between the experimental and predicted water applying the model to flotation columns and used a

18 18
Calculated water flow rate (m /hr)
Calculated water flow rate (m /hr)

16 16
3
3

2 2
14 R = 0.972 14 R = 0.987
12 12
10 10
8 8
6 6
Qa = 1085 l/min Qa = 1085 l/min
4 Qa = 1380 l/min 4 Qa = 1380 l/min
Qa = 1674 l/min Qa = 1674 l/min
2 2 Qa = 1968 l/min
Qa = 1968 l/min
0 0
0 2 4 6 8 10 12 14 16 18 0 2 4 6 8 10 12 14 16 18
3
(a) Experimental water flow rate (m3 /hr) (b) Experimental water flow rate (m /hr)

Fig. 1. Experimental versus fitted concentrate water flow rates using the water–solid model (Eq. (3)) for the Outokumpu 3 m3 tank cell data at
Xstrata Mt Isa copper concentrator: (a) fitted with all data, (b) fitted according to air rate.
874 X. Zheng et al. / Minerals Engineering 19 (2006) 871–882

35 35

Calculated water flow rate (ml/sec)

Calculated water flow rate (ml/sec)


30 2 30 2
R = 0.838 R = 0.868
2 2
25 S = 198.09 25 S = 166.94

20 20

15 15

10 Rougher 1 10 Rougher cell 1


Rougher 2
Rougher cell 2
5 Rougher 3 5
Rougher cell 3
0 0
0 5 10 15 20 25 30 35 0 5 10 15 20 25 30 35
(a) Experimental water flow rate (ml/sec) (b) Experimental water flow rate (ml/sec)

Fig. 2. Experimental versus fitted concentrate water flow rates using the water–solid model (Eq. (3)) for platinum pilot plant data for Harris (2000):
(a) fitted for all cells, (b) fitted for individual cells.

statistical technique to determine the main factors affect- where sf, mean froth residence time; Jg, superficial gas
ing the model results. Three more operating variables velocity entering the froth phase; eg, mean air hold-up
were incorporated in the modified model, including in the froth; Qs, slurry flow rate of the concentrate; Vf,
froth height, wash water and column size: effective volume of the froth zone (the volume occupied
by the moving froth towards the concentrate launder).
J w ¼ ða  Rbs þ c  J b Þ  H df  De ð4Þ Eq. (5) is a simplified version of the following equa-
tion, assuming that the horizontal cross-sectional area
where Jw, superficial water velocity to the concentrate; of the froth zone does not change with level and that
Rs, fractional solids recovery to the concentrate; Jb, bias the water/solids hold-up in the froth phase is negligible
water velocity; Hf, froth height; D, column diameter; a, compared with the volume occupied by the air:
b, c, d and e, constants. eg  V f
It may be useful to extend the model modification of sf ¼ ð7Þ
Qa
Uribe et al. (1999) to mechanical cells, if the interpreta-
tion holds that froth height and flotation cell size (froth where Qa is the rate of air flow passing through the froth
transportation distance) can affect the relationship be- zone.
tween the concentrate water flow rate and the concen- It should be noted that the residence time of the air
trate solids flow rate. bubbles and attached particles in the froth phase usually
In summary, the practical application of the concen- differs from that of the water and entrained particles in
trate water flow rate model as a function of the solids the same froth but the two residence times are closely re-
flow rate given in Eq. (3) is limited to a narrow range lated. It should also be noted that, in industrial flotation
of cell operating conditions and to a narrow range of cells, there is always a significant portion of the froth
flotation cells operated with similar conditions, as the burst on the surface of the cell and not reporting to
froth properties and the solids content in the concen- the concentrate launder (Barian et al., 2005; Zheng
trate depend strongly on the cell operating conditions et al., 2005). As a result, the froth movement slows down
and duties. and the residence time of the froth increases (Zheng and
Knopjes, 2004). Therefore, Eqs. (5) and (7) should be
3.2. Water recovery as a function of froth modified to more accurately reflect the true value of
residence time the mean froth residence time by taking into account
the air recovery factor
Froth residence time is possibly the most frequently eg  V f
used variable in the modelling of froth recovery. Two sf ¼ ð8Þ
a  Qa
definitions of mean froth residence time, based on the
air rate and concentrate slurry flow rate, respectively, where a is the fraction of the air leaving cell as unburst
can be found in the literature (Gorain et al., 1998; Lynch bubbles.
et al., 1981): The main motivation of using Eq. (5) to estimate the
mean froth residence time in flotation modelling is that
Hf
sf ¼ ð5Þ Eq. (5) includes two cell operating conditions only, froth
Jg
height and air rate, which can be either set or measured
ð1  eg Þ  V f easily. It is important to point out that under normal
sf ¼ ð6Þ
Qs operating conditions, the top of the froth is above the
X. Zheng et al. / Minerals Engineering 19 (2006) 871–882 875

level of the concentrate launder weir. Until now, there is mean froth residence time to determine the water recov-
no information available in the literature on how the ery is inadequate to distinguish the effect of air rate and
height of the froth above the concentrate weir level is re- froth height. Based on the experimental data shown in
lated to the cell operating conditions or what is the typ- Fig. 3, it can also be found that water recovery is more
ical range of the froth height above the concentrate sensitive to froth height than to air rate. Neethling et al.
launder lip in different flotation cells and/or when treat- (2000) demonstrated that water distributes unevenly in a
ing different ores. In the rougher test work conducted flowing foam as the froth residence time varies from
using the Outokumpu 3 m3 tank cell in the Xstrata Mt location to location and the properties of the local froth
Isa copper concentrator (Table 1), the height of the froth depends strongly on its residence time. A single value of
above the concentrate launder lip level varied between mean froth residence time does not adequately represent
3 cm and 4.5 cm while the pulp level was set to give a the true situation of froth transportation and cannot be
froth height between 6.2 cm and 20.2 cm (distance be- used to determine the properties (i.e., water content) of
tween the pulp–froth interface and the concentrate laun- the froth entering the concentrate launder. Based on the
der lip). Generally, the froth is higher above the plant observations, the above phenomena cannot be ex-
concentrate launder lip at a deeper froth height setting. plained without carefully examining the change in froth
Therefore, care should be taken when using Eq. (5) to characteristics, especially bubble size and froth stability,
predict the mean froth residence time at a shallow froth at different air rates and froth heights.
and a low air rate. It should also be noted that the water recovery pro-
Savassi (1998) proposed an empirical model for water posed by Savassi (1998) identified the froth residence
recovery as a power function of the mean froth residence time is the key factor controlling the water recovery. In
time (calculated by Eq. (5)) based on the data from his fact, there are two distinguished steps in the water recov-
test work in the lead rougher circuit at the Xstrata Mt ery process, transfer of the water from the pulp phase to
Isa lead/zinc concentrator. the froth phase and transfer of the water in the froth
Rw ¼ c  sdf ð9Þ phase to the concentrate launder. The recovery of the
water from the pulp phase to the froth phase is relatively
where Rw, fractional recovery of the feed water in the
independent of the events in the froth phase but depen-
concentrate; sf, mean froth residence time calculated
dent on the air rate, bubble size distribution in the pulp
using Eq. (5); c and d are constant.
phase, bubble packing condition at the pulp–froth inter-
Fig. 3 plots the experimental water recoveries at dif-
face, and mineralisation of the bubble surface.
ferent cell operating conditions using the data obtained
Meanwhile, Gorain et al. (1998) also used the mean
during the Outokumpu 3 m3 tank cell test work at the
froth residence time in their model but to determine
Xstrata Mt Isa copper concentrator. Eq. (9) was used
the froth recovery only. Froth recovery is expressed as
to fit the experimental data for different air rates
an exponential function of the mean froth residence
(Fig. 3(a)) and for different froth heights (Fig. 3(b)).
time:
Eq. (9) appears to fit the experimental data well both
at different froth heights at a given air flow rate and at Rf ¼ ebsf ð10Þ
different air rates at a given froth height. However, the
water recovery did not fall on the same curve at the dif- where Rf, froth recovery; b, constant, related to the froth
ferent air rates and different froth heights even at the characteristics.
same froth residence time, especially at the lower air The overall recovery of the water is a result of the
rates and shallower froth. Clearly, Eq. (9) using the recovery of the water from the pulp phase to the froth

45 45
40 Qa = 1085 l/min Hf = 6.2-7.2 cm
Qa = 1380 l/min 40
Hf = 10.4-11.8 cm
35 Qa = 1674 l/min Hf = 14.6-16.0 cm
35
Water recovery (%)

Water recovery (%)

Qa = 1968 l/min Hf = 18.8-20.2 cm


30 30
25 25
20 20
15 15
10 10
5 5
0 0
0 5 10 15 20 25 0 5 10 15 20 25
(a) Mean froth residence time (sec) (b) Mean froth residence time (sec)

Fig. 3. Modelling of water recovery as a power function of mean froth residence time (Outokumpu 3 m3 tank cell data at Xstrata Mt Isa copper
concentrator): (a) different air rates, (b) different froth heights.
876 X. Zheng et al. / Minerals Engineering 19 (2006) 871–882

45 60
40 Qa = 1085 l/min Hf = 6.2-7.2 cm
Qa = 1380 l/min 50 Hf = 10.4-11.8 cm
35 Qa = 1674 l/min Hf = 14.6-16.0 cm
Water recovery (%)
Qa = 1968 l/min

Water recovery (%)


Hf = 18.8-20.2 cm
30 40
25
30
20
15 20
10
10
5
0 0
0 5 10 15 20 25 0 5 10 15 20 25
(a) Mean froth residence time (sec) (b) Mean froth residence time (sec)

Fig. 4. Modelling of water recovery as an exponential function of mean froth residence time (Outokumpu 3 m3 tank cell data at Xstrata Mt Isa
copper concentrator): (a) different air rates, (b) different froth heights.

phase and the recovery of the water from the froth phase Figs. 3 and 4 that air rate and froth height affect the
to the concentrate. The water returning from the froth water recovery in a different way and their effect on
phase to the pulp phase may be considered as an addi- the water recovery cannot simply be represented by a
tional feed to the flotation cell pulp. Hence, the overall single value of mean froth residence time.
water recovery can be expressed by
Rcw  Rfw 3.3. First-order water recovery model (Harris, 2000)
Rw ¼ ð11Þ
1  Rcw þ Rcw  Rfw
Harris (2000) proposed a water recovery model which
where Rcw, recovery of the water from the pulp phase to was built on the work of Gorain et al. (1998) and Harris
the froth phase; Rfw, recovery of the water from the (1998). In his model, water recovery from the pulp phase
froth phase to the concentrate, or froth recovery of the to the froth phase and from the froth phase to the con-
water. centrate launder are modelled separately. Water in the
Combining Eqs. (10) and (11) gives: pulp phase is treated as a component in the same way
Rw Rcw as the minerals in the system. Recovery of the water
¼  expðb  sf Þ ð12Þ from the pulp phase to the froth phase is assumed to fol-
1  Rw 1  Rcw
low a first-order kinetic equation (Scrimgeour et al.,
By plotting (Rw/(1  Rw)) against sf, the values of Rcw 1970). Hence, if the pulp phase is considered to be per-
and b can be determined. Then, Rfw can be calculated fectly mixed, then the water recovery can be expressed
from the known b value. Fig. 4 shows the fitting using via Eq. (14):
Eq. (12).
In comparison with the power function model (Eq. kw  s
Rw ¼ ð14Þ
(9)), the exponential model (Eq. (12)) appears also to 1 þ kw  s
fit the experimental data well. However, the same limita- where s, mean residence time of the pulp slurry; kw, first-
tion is still associated with the model by Gorain et al. order rate constant for water, as a function of bubble
(1998) as the one by Savassi (1998) due to use of a single surface area flux in the pulp phase, and
value of mean froth residence to represent the entire
k w ¼ P w  S b  Rfw ð15Þ
froth transport behaviour in the cell, even though the
approach of Gorain et al. (1998) to modelling the water where Pw, constant for a given flotation system; Sb, bub-
recovery does distinguish the two steps of water transfer ble surface area flux in the pulp phase, calculated from
in the flotation cell. the superficial gas velocity and the bubble size.
Finally, it is worth mentioning that applying a simple The model for the froth recovery of water is ex-
exponential function (Eq. (13)) would give a similar fit- pressed as an exponential function of the mean froth res-
ting result. Since the model of Gorain et al. (1998) is an idence time. Two forms of froth recovery models are
empirical one, such a simplification is acceptable in adopted by Harris (2000):
practice. Rfw ¼ ð1  aw Þ  expðbw  sf Þ þ aw ð16Þ
Rw ¼ Rcw  expðb  sf Þ ð13Þ Rfw ¼ X  expðr  V f  v  sf Þ ð17Þ
In summary, it can be concluded that the mean froth where aw, bw, X, r and v are constant.
residence time is an important parameter for determin- Eq. (16) was first proposed by Harris (1998). aw is
ing the froth recovery of the water and the overall water given a physical meaning as the non-draining fraction
recovery. Either power or exponential model fits the of the froth. The modified form (Eq. (17)) excludes the
experimental data well. However, it is clearly shown in non-draining fraction factor but introduces a new coef-
X. Zheng et al. / Minerals Engineering 19 (2006) 871–882 877

ficient related to the fraction of stagnant plus collapsed recovery but also the effect of froth volume. The addi-
froth (X). tional parameter in the model allows the effects of air
Combining Eqs. (14)–(16), the overall water recovery rate and froth height to be considered and modelled sep-
can also be expressed by arately (to a certain degree). This is important especially
in large industrial flotation cells where the froth rheolog-
1
Rw ¼ 1  ð18Þ ical behaviour may differ significantly with location and
1 þ P w  S b  s  X  expðr  V f  v  sf Þ stagnant froth may develop in certain areas of the froth
The above model was reported to fit the same pilot plant phase.
data (as shown in Fig. 2) well, and be able to predict the In summary, the approach of Harris (2000) to model-
water recovery with a reasonable accuracy (Harris, ling the recovery of the water from the pulp phase to
2000). However, if comparing the goodness of fitting the froth phase as a first-order process and modelling
using the Harris (2000) approach (Fig. 5(a)) with the fit- the froth recovery of the water as a function of both
ting results using Savassi (1998), Gorain et al. (1998) mean froth residence time and froth volume appears
models (Fig. 5(b)) to the same data obtained in the to fit the Xstrata Mt Isa plant data well. However, there
Outokumpu 3 m3 tank cell data at the Xstrata Mt Isa are still questions to be answered:
copper concentrator, it is clear that, despite the attempt
of Harris (2000) to develop the water recovery model a • What is the relationship between the water recovery
step further from Eq. (14) by incorporating a model from the pulp phase to the froth phase and the mean
for the recovery of the water from the pulp phase to pulp residence time?
the froth phase using a first-order process, the results • How to distinguish the constant for water recovery
do not seem to offer any improvement from the simpler from the pulp to the froth (Pw) and the efficiency fac-
models of Savassi (1998) and Gorain et al. (1998). tor for the froth volume and bubble bursting (X) (the
It should be noted that, although the model of water two are indistinguishable in the model, Eq. (18)).
recovery from the pulp phase to the froth phase has the
form of a first-order process, it is not the authors inten- 3.4. Drainage model for water recovery (Moys, 1984)
tion to suggest the water recovery mechanism is indeed
first-order: it is a purely empirical model which attempts In the model of Moys (1984), water recovery is also
to correlate the water recovery with the cell operating considered as a two-step process. The pulp water enters
variables. However, as mentioned at the beginning of the froth phase in the film of the air bubbles. In the froth
the paper, it is still debatable whether the amount of phase, water needs to survive the drainage process and
water entering the froth phase is related to the feed bubble bursting on the surface before it can reach the
water flow rate. The results from the Outokumpu 3 m3 concentrate launder.
tank cell tests at the Xstrata Mt Isa copper concentrator To determine the initial water flow rate from the pulp
with different feed flow rates show a nearly unchanged phase to the froth phase, it is assumed that the water is
concentrate water flow rate at a given air rate and froth carried by the air bubbles and that the film thickness is
height when the cell was operated as the first rougher constant in a given flotation system.
and the bubble loading condition did not change signif-
Qw ð0Þ ¼ A  S b  d ð19Þ
icantly with the different feed rates.
It should also be noted that Eq. (18) takes into ac- where Qw(0), water flow rate entering the froth phase at
count not only the effect of froth residence time on froth the pulp–froth interface; A, cross-sectional area of the

40 40
Calculated water recovery (%)

35 35 Savassi (1998), R2 = 0.9895, S2 = 18.01


2
Gorain et al (1998), R2 = 0.9895, S2 = 16.68
Calculated water recovery

R =0.9663
30 2 30
S =56.91
25 25

20 20
15 15
Qa = 1085 litre/min
10 Qa = 1380 litre/min 10
Qa = 1674 litre/min
5 Qa = 1968 litre/min 5
0
0
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
(a) Experimental water recovery (b) Experimental water recovery (%)

Fig. 5. Comparison of the model fitting results using the Outokumpu 3 m3 tank cell data at the Xstrata Mt Isa copper concentrator: (a) Harris
(2000), (b) Savassi (1998) and Gorain et al. (1998).
878 X. Zheng et al. / Minerals Engineering 19 (2006) 871–882

pulp–froth interface; d, the volume of water per surface both the air bubbles and the slurry and comparing it
area of the air bubbles. with the total air introduced to the cell.
During the transfer process in the froth phase, water Qca
drains back to the pulp phase at a rate proportional to a¼ ð27Þ
Qa
the water concentration at the froth level of interest.
where Qca is the volumetric flow rate of the froth (the
dQw ðzÞ air) in the concentrate.
¼ k wd  Qw ðzÞ ð20Þ
dtf There are only two fitted parameters in Eq. (26)—the
volume of water per surface area of the air bubbles
where tf, time for the froth to rise from the pulp–froth
entering the froth phase and the drainage rate constant
interface to the level of interest in the froth; kwd, drain-
in the froth phase. Fig. 6 shows the results of fitting the
age rate constant of the water.
model to the data from the test work at the Xstrata Mt
Since the froth residence time can be expressed as
Isa copper concentrator using the Outokumpu 3 m3
ef ðzÞ  AðzÞ  dz tank cell. It should be noted that model fitting is carried
dtf ¼ ð21Þ out for one air rate at a time, i.e., the model fitting was
Qa
conducted independently for the different air rates. As
where ef(z), air hold up at the froth level of interest; A(z), can be seen, Moys model (1984) fits the concentrate
cross-sectional area of the froth phase at the level of water flow rate very well.
interest. Fig. 7 further examines the effect of air rate on the
Eq. (20) can be re-written as water recovery from the pulp phase to the froth phase
dQw ðzÞ and the effect of froth residence time on the froth recov-
¼ k wd  ef ðzÞ  AðzÞ  Qw ðzÞ=Qa ð22Þ ery of the water using the same set of data obtained at
dz
the Xstrata Mt Isa copper concentrator.
If the froth cross-sectional area remains constant In general, there is a trend of increase in the water
throughout the entire froth phase and the change in flow into the froth phase from the pulp phase with in-
air hold-up at different froth levels is negligible and crease in the air rate. However, the curve of the pulp
the drainage rate constant is assumed to be independent water recovery to the froth phase and air rate in
of froth level, then Eq. (22) can be integrated from the Fig. 7(a) shows an irregular shape. It is difficult to use
pulp–froth interface as the zero level to the top of the the water film theory to explain this phenomenon.
froth phase, which gives the solution as Clearly, at a higher air rate, there are more air bubbles
Rfw ðH f Þ ¼ expðk wd  ef  A  H f =Qa Þ ð23Þ rising through the pulp–froth interface. If the bubble
size remains the same at different air rates, then the
where Hf is the total froth height. amount of water recovered from the pulp phase should
Because the (ef Æ A Æ Hf/Qa) term can be replaced by be proportional to the air rate. However, in reality,
froth residence time, the froth recovery of the water at water is also carried into the froth phase in the Plateau
the top of the froth phase becomes: borders. The amount of water in the Plateau borders de-
Rfw ðH f Þ ¼ expðk wd  sf Þ ð24Þ pends not only on the bubble size and number but also
the packing condition of the bubbles on the pulp–froth
Not all the froth reaching the top of the froth phase en-
interface. If the bubbles are closely packed at the
ters the concentrate. Some of the bubbles burst on the
surface. If the fraction of the froth on the surface that
eventually reports the concentrate is a, then the froth 16
Calculated concentrate water flow (ton/hr)

recovery of the water becomes: 14

Rfw ¼ a  expðk wd  sf Þ ð25Þ 12 2


R = 0.9807
2
S = 4.39
The froth recovery model given by Eq. (25) is identical 10

with the one proposed by Gorain et al. (1998), which 8


was obtained empirically (Eq. (10)).
6
Combining Eqs. (19) and (25), the concentrate water Qa = 1085 l/min
4
flow rate can be determined: Qa = 1380 l/min
Qa= 1674 l/min
2
Qw ¼ A  S b  d  a  expðk wd  sf Þ ð26Þ Qa = 1968 l/min
0
Generally, the cross-sectional area of the pulp–froth 0 2 4 6 8 10 12 14 16
interface is known. The bubble surface area flux in the Experimental concentrate water flow (ton/hr)

flotation cell can be measured. Moys suggests that the Fig. 6. Experimental versus fitted concentrate water flow rates using
efficiency of froth removal be estimated by measuring Moys model (1984) for the Outokumpu 3 m3 tank cell data from
the volumetric flow rate of the concentrate including Xstrata Mt Isa copper concentrator.
X. Zheng et al. / Minerals Engineering 19 (2006) 871–882 879

60
35
Water flow from pulp to froth (m3/hr) Qa = 1085 l/min

Froth recovery of the water (%)


Qa = 1380 l/min
30 50
Qa = 1674 l/min
-0.0709x
y = 100.85e Qa = 1968 l/min
25 40
2
R = 0.9997

20
30
15
20 -0.1175x
y = 96.101e
2
10 R = 0.9896

10
5

0 0
1000 1200 1400 1600 1800 2000 2200 0 5 10 15 20 25 30
(a) Air rate (l/min) (b) Froth residence time (sec)

Fig. 7. Moys model fitting results on Xstrata Mt Isa copper concentrator test work data: (a) effect of air rate on pulp water recovery, (b) effect of
froth residence time on froth recovery.

pulp–froth interface, the liquid content at the interface is crease in water content is caused predominantly by coa-
about 0.26, while if the packing is random, the value is lescence, i.e. increase in bubble radii and therefore a
about 0.36 (Neethling et al., 2003). Interestingly, the val- decrease in the total length of Plateau borders. Test
ues calculated from the Outokumpu 3 m3 tank cell data work and measurement were conducted both in a two-
were 0.26 and 0.30 at the air rates of 1968 l/min and phase foam in the laboratory and in a three-phase froth
1974 l/min, respectively. in the industrial flotation cell, and a froth recovery
It is also interesting to see that the froth recovery of model for was developed which calculates the length
the water calculated from the model falls on the same of the Plateau borders and the cross-sectional area of
curve at the two high air rates and again at the two the Plateau borders.
low air rates tested in the Outokumpu 3 m3 tank cell The numerical solution for the length of Plateau bor-
(Fig. 7(b)). The possible explanation is that the froth ders per volume of the froth is based on the dodecahe-
structure and properties are similar within the two oper- dron geometry of the froth:
ating air rate ranges. Further investigation is required. pffiffiffi
In summary, Moys model (1984) produces a good fit 5 3 6:81
k¼ or k ¼ 2 ð28Þ
to the experimental data for the overall recovery of the p  u  r2 df
water, but the physical significance of the model
parameters, especially the water film thickness of the where k, length of Plateau borders per volume pof ffiffi the
air bubbles, may not reflect the true water recovery froth; r, bubble radius; u, Golden ratio ð¼ 1þ2 5Þ; df,
mechanisms. bubble diameter.
The cross-sectional area of the Plateau borders in the
3.5. Fundamental approach to modelling of froth froth phase is determined through a balance of forces
recovery of water (Neethling et al., 2003) acting on the water in the Plateau borders. Three main
forces are considered: gravitational forces, capillary
Research into understanding the foam/froth struc- forces and forces due to viscous dissipation. For the
ture, developing measurement techniques and modelling froth on the top of the cell, capillary force may be ne-
the water recovery in flotation has been advanced signif- glected. Hence, the solution for the cross-sectional area
icantly in the recent years, especially by a group of of the Plateau borders on the top of the froth is given as
researchers at the UMIST (Neethling et al., 2000,
2003; Neethling and Cilliers, 2003; Barian et al., 2005). Qa g  150
At ¼  ð1  aÞ  ð29Þ
The research outcomes have now become available for W D qg
applications in the mineral processing industry (Nee-
where At, cross-sectional area of the Plateau borders on
thing and Cilliers, 2005).
top of the froth; W and D, dimensions of the top of the
In brief, Neethling and Cilliers (2003) stated that
cell; a, fraction of the air leaving cell as unburst bubbles;
most of the froth water is held and moves in the Plateau
g, water viscosity in the froth, q, water density; g, grav-
borders, which present in the froth phase as a network
itational constant.

of interconnected, nearly randomly distributed drainage
The WQD a
term is in fact the superficial gas velocity,
channels. In a froth phase that is at steady state and has
Jg. Combining Eqs. (28) and (29) gives the volumetric
a relatively low gas velocity, the cross-sectional area of
fraction of the water at the top of the cell:
the Plateau borders decreases rapidly with height above
the pulp–froth interface before reaching the concentrate 255:6 g
e w ¼ At  k ¼ 2
 J g  ð1  aÞ  ð30Þ
overflow level. For most heights of the froth, any de- r qg
880 X. Zheng et al. / Minerals Engineering 19 (2006) 871–882

where ew is the water hold-up at the top of the froth for the calculation of the length of the Plateau borders.
surface. In addition, a large portion of the froth entering the con-
It should be taken into account that only a fraction of centrate is coming from the froth flow beneath the sur-
the air bubbles remaining unburst and reporting to the face layer. The actual bubble size could be significantly
concentrate launder: smaller than that measured on the surface. Therefore,
Qac ¼ a  Qa ð31Þ for the purposes of model fitting, it is convenient to as-
sume that the size of the surface bubbles increases line-
The a value can be estimated through measurement of arly with increase in the froth height:
the velocity of the froth flowing over the lip, the height
r ¼ r0 þ a  H f ð34Þ
of the froth above the concentrate launder lip and the
volumetric flow rate of the air into the cell: where r0, radius of the air bubble initially entering the
v f  hf  l froth phase; a, constant.
a¼ ð32Þ The measured bubble size on the cell surface is also
Qa
affected by the froth stability. When the froth is unstable
where vf, velocity of the froth flowing over the lip; hf, and bubble bursting occurs rapidly, fine bubbles are ex-
froth height above the concentrate launder lip; l, the posed between the large bubbles. This often happens in
total lip length of the concentrate launder. a high air rate or at a low bubble loading condition.
Hence, the concentrate water flow rate can be Fig. 9 is the image of the surface froth at four different
obtained as heights (increase from the top image to the bottom
ew one) at two air rates tested in the Outokumpu 3 m3 tank
Qw ¼ a  Qa  cell.
1  ew
! Accurate measurement of the air recovery is also dif-
1 ficult. To collect the unburst bubbles in the concentrate
¼ a  Qa  g 1 ð33Þ
1  255:6
r2
 J g  ð1  aÞ  qg launder is almost impossible. The image processing
method (Eq. (32)) depends strongly on how representa-
It should be noted that the above equation ignores the
tive the measured froth velocity and froth height above
water in the lamellae and takes no account of the capil-
the concentrate launder lip are of the average values
lary term. In addition, variation in the froth properties
(Neethling et al., 2003). Hence, again for modelling pur-
across the horizontal direction, i.e., bubble size and
poses, the fraction of unburst bubbles on the surface of
froth viscosity, is also ignored. The key parameters are
the cell is considered to be linearly proportional to the
the bubble size and the fractional recovery of the air
froth height:
bubbles in the concentrate launder. Fig. 8 shows the
measured bubble size at the different froth heights at
the different air rates.
In general, the size of the bubbles on the cell surface
increased with increase in the froth height in an approx-
imately linear manner. However, it should be noted that
using image analysis to determine the surface bubble size
in two dimensions may not truly represent the bubble
shape in three dimensions, which is especially important

14

12
Bubble Diameter (mm)

10

4 Qa =1085 l/min
Qa = 1380 l/min
2 Qa = 1674 l/min
Qa = 1968 l/min

0
0 5 10 15 20 25 30
Froth Height (cm)
Fig. 9. Froth images at different froth heights and air rates in the
Fig. 8. Measured mean bubble diameter at different froth heights at Outokumpu 3 m3 tank cell at the Xstrata Mt Isa copper concentrator
different air rates in the Outokumpu 3 m3 tank cell data at the Xstrata (same scale for all the images): (a) air rate = 1085 l/min, (b) air
Mt Isa copper concentrator. rate = 1674 l/min.
X. Zheng et al. / Minerals Engineering 19 (2006) 871–882 881

use of a large froth crowder (1 m in diameter) in the test


Calculated concentrate water flow rate (m3/hr)

16

14
cell which has a diameter of 1.3 m at the top of the cell.
The most significant feature of the fundamental
12
2
model (Eq. (33)) is that all the variables in the model
R = 0.9668
10 2
S = 7.82 have physical meaning. They can be determined experi-
mentally, i.e. a and r from a froth monitoring device,
8
and Qa and Jg from an air flow meter. Hence, the model
6 is ready for use in practice for process control purposes.
4 Another advantage of the model (Neethling et al., 2003)
over the other empirical models in this paper is that it
2
fits all the experimental data well at the same time
0 regardless of the cell operating conditions. By incorpo-
0 2 4 6 8 10 12 14 16
rating additional models to relate the air recovery and
Experimental concentrate water flow rate (m3/hr)
bubble size on the surface to the cell operating condi-
Fig. 10. Experimental versus fitted concentrate water flow rates using tions directly, the fundamental model will also be able
Neethling et al.s model (2003) based on the data obtained in the to use for prediction purposes.
Outokumpu 3 m3 tank cell at the Xstrata Mt Isa copper concentrator.

Hf 4. Summary
a¼1 ð35Þ
H f ðmaxÞ
No matter whether they are empirically or fundamen-
where Hf (max) is the maximum froth height that can be
tally based, the mathematical models for water recovery
obtained under the given operating conditions. At this
or concentrate water flow rate reviewed in this paper
maximum froth height, all bubbles burst.
tend to fit the experimental data reasonably well for a
Fig. 10 shows the fitted water recovery using Eqs.
given flotation system. However, the empirical models
(33)–(35) in comparison with the experimentally mea-
which relate the water recovery to the solids recovery
sured values.
(Alford, 1990) or describe the water recovery as a func-
The fitted results using the model developed by
tion of the mean froth residence time (Savassi, 1998) are
Neethling et al. (2003) match the experimental ones sat-
generally unable to distinguish the effect of different cell
isfactorily. Fig. 11 compares the measured bubble size
operating conditions such as froth height and air rate.
on the surface with the fitted value.
Hence, these models can only be used within a relatively
Although there is a significant difference between the
narrow range of cell operating conditions. The empirical
measured bubble size and the model fitted one, there is
models which consider the water recovery as a two-step
clearly a strong correlation between them. It is an area
transfer (Gorain et al., 1998; Harris, 2000) may be used
in the future to improve the measurement technique
for prediction purposes if the model parameters can be
and/or develop a model for calibration.
derived beforehand. However, often these parameters
The air recovery calculated is between 25% and 71%
have no physical meaning and have to be derived in
in the Outokumpu 3 m3 tank cell, higher than the values
an existing system. The fundamental approach adopted
reported for the other operations (Barian et al., 2005;
by Neethling et al. (2003) resulted in a water recovery
Zheng et al., 2005). This may attribute largely to the
model with all the variables directly related to the phys-
ical conditions in the flotation process. The variables can
0.45
also be determined experimentally. Hence, the model is
0.40 ready for use in practice for process control purposes.
Calculated bubble size (cm)

0.35 Once the bubble size on the surface and the air recovery
0.30 in the flotation system can be modelled, the water recov-
0.25 ery model developed by Neethling et al. (2003) can also
0.20 be used for prediction purposes.
0.15
Qa =1085 l/min
0.10 Qa = 1380 l/min

0.05
Qa = 1674 l/min Acknowledgements
Qa = 1968 l/min

0.00
0 0.2 0.4 0.6 0.8 1 1.2 1.4 The authors would like to thank the sponsors of the
Measured bubble size (cm) AMIRA P9 project for the funding which made this
Fig. 11. Comparison of the measured and model fitted bubble size on work possible. The authors would also like to thank
the cell surface based on the data obtained in the Outokumpu 3 m3 Outokumpu for the loan of the 3 m3 tank cell used in
tank cell at the Xstrata Mt Isa copper concentrator. the test work. Support from the staff at the Xstrata Mt
882 X. Zheng et al. / Minerals Engineering 19 (2006) 871–882

Isa copper concentrator, especially Mr. David Carr, dur- Neethling, S.J., Cilliers, J.J, Woodburn, E.T., 2000. Prediction of the
ing the on-site test work is also gratefully acknowledged. water distribution in a flowing foam. Chemical Engineering Science
556, 4021–4028.
Neethling, S.J., Lee, H.T., Cilliers, J.J., 2003. Simple relationships for
predicting the recovery of liquid from flowing foams and froths.
References Minerals Engineering 16, 1123–1130.
Neethling, S.J., Cilliers, J.J., 2003. Modelling flotation froths. Inter-
Alford, R.A., 1990. An improved model for design of industrial national Journal of Mineral Processing 72, 267–287.
column flotation circuits in sulphide applications. In: Gary, P.M.J. Neething, S.J., Cilliers, J.J., 2005. The simulation of flotation banks
(Ed.), Sulphide Deposits—Their Origin and Processing. IMM, and circuits, Proceedings Centenary of Flotation Symposium,
London. Brisbane, Australia.
Barian, N., Hadler, K., Ventura-Medina, E., Cilliers, J.J., 2005. The Savassi, O.N., 1998. Direct estimation of the degree of entrain-
froth stability column: linking froth stability and flotation perfor- ment and the froth recovery of attached particles in industrial
mance. Mineral Engineering 18, 317–324. flotation cell, Ph.D. thesis (unpublished), University of Queens-
Gorain, B.K., Harris, M.C., Franzidis, J.-P., Manlapig, E.V., 1998. land.
The effect of froth residence time on the kinetics of flotation. Scrimgeour, J.H.C., Hamilton, R.E., Toong, T., 1970. The use of
Minerals Engineering 11 (7), 627–638. mathematical modeling in developing advanced control system for
Harris, M.C., 1998. The use of flotation plant data to simulate mining industry processes. Transactions of the Canadian Institute
flotation circuits, SAIMM Conference. of Mining 73, 305–314.
Harris, T.A., 2000. The development of a flotation simulation and Uribe, A.S., Vazuez, D.V., Perez, R.G., Nava, F.A., 1999. A statistical
methodology towards an optimisation study of UG2 platinum model for the concentrate water in flotation columns. Minerals
flotation circuits, Ph.D. thesis (unpublished), University of Cape Engineering 12 (8), 937–948.
Town. Zheng, X., Knopjes, L., 2004. Modelling of froth transportation in
King, R.P., 1973. Model for the design and control of flotation plants. industrial flotation cells Part II: Modelling of froth transportation
Society of Mining Engineering of AIME Transactions 274, 341–350. in an Outokumpu tank flotation cell at the Anglo Platinum
Lynch, A.J., Johnson, N.W., Manlapig, E.V., Thorne, C.G., 1981. Bafokeng—Rasimone Platinum Mine (BRPM) concentrator. Min-
Mineral and coal flotation circuits—Kinetic models, Developments eral Engineering 17/9-10, 989–1000.
in Mineral Processing, vol. 3, pp. 64–96. Zheng, X., Franzidis, J.P., Johnson, N.W., Manlapig, E.V., 2005.
Moys, M.H., 1984. Residence time distribution and mass transport in Modelling of froth transportation in an Outokumpu 3 m3 tank cell,
the froth phase of the flotation process. International Journal of Proceedings Centenary of Flotation Symposium, Brisbane,
Mineral Processing 13, 117–142. Australia.

You might also like