You are on page 1of 10

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/303826132

The effect of cell hydrodynamics on flotation kinetics

Thesis · December 2012

CITATIONS READS

6 441

1 author:

Erico Tabosa
Hatch
37 PUBLICATIONS 195 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

The effect of cell hydrodynamics on flotation View project

All content following this page was uploaded by Erico Tabosa on 23 January 2019.

The user has requested enhancement of the downloaded file.


International Journal of Mineral Processing 156 (2016) 99–107

Contents lists available at ScienceDirect

International Journal of Mineral Processing

journal homepage: www.elsevier.com/locate/ijminpro

The effect of cell hydrodynamics on flotation performance


Erico Tabosa ⁎, Kym Runge, Peter Holtham
Metso — Process Technology and Innovation, 1 Technology Court, Pullenvale, Queensland 4069, Australia

a r t i c l e i n f o a b s t r a c t

Article history: It is clear that along with gas dispersion characteristics, the energy dissipated by the impeller is process determin-
Received 15 February 2016 ing in flotation, and its effect on flotation kinetics has been widely studied. However, turbulent conditions inside a
Received in revised form 27 May 2016 flotation cell have usually only been changed by varying impeller speed or air flowrate or both. Therefore, there is
Accepted 30 May 2016
a need to investigate not only these variables but also how changes in impeller/stator mechanism size and design,
Available online 1 June 2016
cell aspect ratio and cell design affect turbulence. This should lead to a better understanding of the effect of cell
Keywords:
hydrodynamics on flotation performance. The aim of this work was to evaluate the role of flotation cell hydrody-
Flotation cell hydrodynamics namics on flotation performance in a fully instrumented 3 m3 cell. The cell was operated at a copper concentrator
Collection zone flotation rate in Australia with different combinations of airflow rates, impeller speeds and sizes and cell aspect ratio providing
Turbulence a wide range of hydrodynamic conditions. An analysis of the flotation cell performance showed that the overall
Froth recovery copper recoveries were very similar for the conditions tested. However, by decoupling pulp effects from froth ef-
Flotation cell aspect ratio fects it was possible to determine whether the changes made affected the pulp zone and/or froth zone responses.
The analysis showed that the overall recovery had the potential to be up to 10% higher if not limited by froth re-
covery. Comparing the metallurgical performance of the cell with the different hydrodynamic conditions it was
found that the collection zone flotation rate was not directly related to overall energy dissipation, as is commonly
observed at laboratory scale, where energy is usually only changed by varying impeller tip speed. Results suggest
that it is the size of the turbulent zone, rather than just energy input, that affects flotation recovery.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction quiescent zone is believed to be important for flotation as it minimises


the detrimental effects of turbulence on the froth phase.
Three distinct zones can be identified within a mechanical flotation Turbulence in mechanical flotation cells has been quantified by im-
cell: the turbulent, quiescent and froth zones. It is known that the highly peller speed or power input by many researchers (Ahmed and Jameson,
turbulent zone is responsible for the key sub-processes (Schubert and 1985; Deglon et al., 2000; Schubert and Bischofberger, 1979; Schubert,
Bischofberger, 1978, 1979; Schubert et al., 1982; Schubert, 1985, 1986, 1985, 2008; and many others), but few have considered the parameters
1999, 2008; Weiss and Schubert, 1988; Weber et al., 1999): necessary for a more comprehensive characterisation of turbulence.
Exceptions are Jordan and Spears (1990); Pyke et al. (2002, 2003)
• Suspension of the particles in the pulp; and Duan et al. (2003) who measured turbulence parameters such as
• Feeding and dispersion of the air into bubbles; turbulent fluctuating velocities and energy spectra in batch flotation
• Mixing of the aerated pulp for better reagent distribution and condi- cells and incorporated these results into micro-kinetic models for flota-
tioning; and tion in agitated systems. The authors considered the turbulent microen-
• Promotion of particle-bubble collisions (the basic requirement for vironment in the flotation cell to have a significant impact on both
particle-bubble attachment). bubble breakup and bubble-particle contact.
Schubert and Bischofberger (1978) found that an increase in recov-
Too much turbulence can have an adverse effect on flotation because ery occurred with increasing impeller speed and hence established op-
it results in the detachment of particles from bubbles as they rise in the timum power inputs for different particle size fractions in a cassiterite
pulp, as well as causing destabilisation and loss of recovery within the flotation plant. An optimum in the flotation rate as a function of impeller
froth phase. Thus, a quiescent zone, less energy intensive than the tur- speed, particle size, density, and bubble size was found by Ahmed and
bulent zone, provides conditions for detaching entrained or entrapped Jameson (1985).
gangue particles from aggregates created in the turbulent zone. The Other similar studies, showing an increase in flotation rate of fine
particles with an increase in impeller speed, can be found in the litera-
ture for different ore types: galena (Spears and Jordan, 1989), chalcopy-
⁎ Corresponding author. rite (Jordan and Spears, 1990) and coal (Cheng et al., 1995).

http://dx.doi.org/10.1016/j.minpro.2016.05.019
0301-7516/© 2016 Elsevier B.V. All rights reserved.
100 E. Tabosa et al. / International Journal of Mineral Processing 156 (2016) 99–107

The energy dissipation, or specific power input (ε), may be consid- ratio providing a wide range of hydrodynamic conditions. The objective
ered equal to the average energy dissipation in a flotation machine was to gain a better understanding of how flotation cell operational var-
(Schubert, 1986, 2008), which is related to the power input (P) of the iables, which change both the energy inputted to the flotation cell and
impeller into the total mass of the fluid system (m) by: how this energy is distributed, impact flotation recovery.

P 2. Methodology
ε¼ ð1Þ
m
2.1. The experimental rig and test work program
This equation shows that achieving high local energy dissipation
rates in the impeller stream is dependent on the power input to the im- The experimental work was carried out in Metso's flotation test rig
peller. The degree of turbulence is at a maximum in the impeller region at the Northparkes copper/gold concentrator, located in the central
and is significantly reduced in the quiescent zone above the impeller west of New South Wales, Australia. The test rig consists of two modules
(Wu and Patterson, 1989; Rutherford et al., 1996; Pyke et al., 2002, that can be disassembled to facilitate transportation (Fig. 1). A 3 m3
2003; Newell and Grano, 2007). Schubert and Bischofberger (1979) Metso RCS flotation cell is located on the upper platform, with concen-
suggest that, in the highly turbulent zones of the cell, values of ε may trate and tailings being discharged via gravity to a product sump tank
be 5 to 30 times higher than the average value. located on the bottom platform. The staircase and the grid mesh plat-
The impeller-stator system area and its immediate vicinity represent form enable personnel to safely access the top of the cell, where differ-
the high intensity zone, where air is sheared and bubbles are generated, ent measurements can be carried out in the flotation cell.
and significant turbulence and air void fraction are present. This zone The 170 cm diameter RCS flotation cell incorporated a variable side-
also represents the area where particles and bubbles come into first wall height, allowing comparison of a standard aspect ratio cell (145 cm
contact. These factors suggest this zone presents the highest probability height), with a lower aspect ratio cell (110 cm height, 2.5 m3 volume).
of collision, and thus the highest flotation kinetics (Koh et al., 2000). By reducing the cell height, the ratio between high and low shear
Hence, high collision rates necessitate the use of an efficient impeller- zones in the cell was increased. The cell geometries are illustrated in
stator system (Schubert, 2008). Fig. 2.
The effect of the high intensity turbulence on flotation recovery has One of the main advantages of the Metso rig is that it incorporates
been widely studied by Schubert and co-workers (Schubert et al., 1982; easily accessible full stream sampling points and is fully instrumented
Schubert and Bischofberger, 1998; Schubert, 1999; and others). They in- to allow stable operation, essential requirements for metallurgical test
vestigated recovery in the high intensity zone by means of high-speed work (Runge et al., 2007).
photography, showing that particle-bubble collection occurs almost The rig can be operated in three modes: open circuit, full recycle and
solely in this zone of high local energy dissipation. tailing recycle (Runge et al., 2007). The configuration used for this test
It is clear that along with gas dispersion characteristics, the energy work was full recycle mode, in which the concentrate and tailing
dissipated by the impeller is process-determining in flotation. Its effect streams are mixed in the product sump and then recycled back as
on flotation kinetics has been widely studied (Schubert and new feed. In this case the cell operates with nominally unchanging
Bischofberger, 1978; Jameson and Ahmed, 1983; Deglon et al., 1999; feed properties. The measured feed flow is kept constant by controlling
Pyke et al., 2003; Newell and Grano, 2006). the product sump pump speed. Feed flow rate was maintained at
Conventional mechanical flotation cell size has increased significant- 80 m3 h−1 throughout the test work, resulting in a flotation residence
ly over the last 20 years. However, as cell size has increased, most man- time of 2.3 min which is typical of that observed in industrial flotation
ufacturers have kept the aspect ratio of the cell constant to maintain cells. A peripheral launder with froth crowder was used to collect the
geometric consistency. This has resulted in cells with a highly turbulent froth.
region near the impeller and very large quiescent zones, which poten- Slurry to fill the test rig was taken from the flotation feed of Module
tially result in very little flotation recovery. There may be an opportunity 1 of the Northparkes concentrator. Flotation feed typically assayed 0.3%
to increase cell recovery rates by changing the volume ratio between Cu at approximately 30% solids by mass, with a P80 of 100 μm. When
the high and low turbulent zones in the cell. processing the flotation feed material, the reagent suite added to the
Most of the studies performed to investigate the effect of energy on test rig was identical to that used in the plant (PAX, frother, NaHS and
flotation response have usually only involved changing impeller speed promoter).
and/or air rate. Only limited laboratory scale work has been reported At the start of each day the rig was filled with fresh slurry from the
in the literature evaluating the size of the high turbulent zone and the plant. The rig was then operated at a series of different test conditions
ratio between high and low turbulent (or quiescent) zones on flotation and measurements performed. A factorial design was used in which
performance. Earlier studies (Hall and Dell, 1989; Morris and each of the following variables was tested at two or three levels:
Matthesius, 1988) on the effect of the ratio of quiescent to turbulent
• Cell aspect ratio (height:diameter): 0.85 and 0.65;
zone volumes in batch tests concluded that turbulence appears to play
• Impeller design: RCS3 (standard impeller for a 3 m3 Metso RCS flota-
an important role in promoting particle-bubble attachment, as well as
tion cell) and RCS5 (oversized impeller, i.e. impeller for a 5 m3 Metso
the generally accepted roles of air diffusion and suspension of solids.
RCS flotation cell);
Both studies suggested that the ratio between high and low turbu-
• Impeller tip speed: 3.5 m s−1 and 6.2 m s−1;
lent zones within a flotation cell affects performance. Hall and Dell
• Air flowrate: 100 and 140 m3 h−1
(1989) showed that in a cell with a larger turbulent zone, i.e. a shorter
• Froth depth: three different froth depths, (typically 5, 9 and 13 cm).
cell, and at a certain optimum air rate, higher flotation rates were ob-
tained. However, Morris and Matthesius (1988) found that for coarser A total of 103 test conditions were evaluated during the test cam-
particles (N 150 μm) the beneficiation efficiency was enhanced when paign. This included the 48 test conditions of every combination of the
flotation was carried out in a cell double in height. factor levels in a full factorial experimental design; and also reproduc-
There would therefore seem to be opportunities to improve flotation ibility testing.
recovery rate by optimising the distribution of turbulence in a flotation To change between aspect ratios, the current sidewall was unbolted,
cell. The work described in this paper was performed to evaluate the removed with a crane and replaced with a new sidewall. To change
role of flotation cell hydrodynamics on flotation performance in a fully between impellers, the current impeller and stator had to be unbolt-
instrumented 3 m3 cell. The cell was operated with different combina- ed and the new combination bolted into place. These modifications
tions of airflow rates, impeller tip speeds and sizes and cell aspect were time consuming and difficult, logistically and therefore there
E. Tabosa et al. / International Journal of Mineral Processing 156 (2016) 99–107 101

Fig. 1. Metso 3 m3 test rig installed at Northparkes copper/gold concentrator.

was no choice but to perform all testwork associated with a particu- • Metallurgical performance (feed, concentrate and tail copper grades);
lar impeller/aspect ratio in a block before moving onto a new combi- • Volumetric flowrates (feed and concentrate);
nation. Randomisation of the test sequence was therefore not • Overall copper recovery (Ro);
possible. • Overall flotation rate (k);
Each day of test work, six conditions were evaluated for a particular • Collection zone flotation rate (kc);
impeller/aspect ratio combination. A particular air rate, impeller tip • Froth recovery (Rf);
speed combination would be randomly chosen and then the pulp level • Collection zone recovery (Rc); and
(froth depth) would be set to a shallow, intermediate and deep froth • Power input to the impeller and overall energy dissipation
depth. Another combination of impeller tip speed and air flowrate was characterisation.
then set and tests were again run for different froth depths. The aim
was to perform tests at different froth depths for a particular combina- To enable the concentrate grade and recovery to be determined,
tion of variables on the same day over a relatively short time interval samples of the feed to the flotation cell, concentrate and tailings were
to maximise the accuracy of the technique used to estimate collection collected. All samples were weighed wet, filtered and dried, weighed
and froth zone recoveries (outlined below). dry and analysed for copper in the Northparkes laboratory.
Besides the six conditions tested per day, a standard condition was A mass balance was performed for each test condition to establish
set at the beginning of each day in order to assess the day-to-day repro- unmeasurable flows and to make all measurements consistent. This
ducibility of the tests. The standard condition was 3.5 m s−1 impeller tip was done by using the measured feed, concentrate and tails assays
speed, 100 m3 h−1 air flowrate and 9 cm froth depth. This condition was and solids concentration data as well as the measured concentrate and
also repeated six times during one testing day to assess reproducibility feed flowrates. This balance was set up in a Microsoft Excel spreadsheet
of the testing procedure over a day of operation. with Solver used to find the flows/assays that resulted in a minimum
The main response variables measured at each condition were as sum of squares error between the experimental data and the mass bal-
follows: anced data.

Fig. 2. (a) High and (b) low cell aspect ratio configurations.
102 E. Tabosa et al. / International Journal of Mineral Processing 156 (2016) 99–107

2.2. Analysis Vera et al. (1999) showed that froth recovery can be defined as the
total rate of transfer from the pulp to the concentrate (k) divided by
Using the mass balance data, the overall copper recovery (Ro) was the rate of transfer from the pulp to the froth phase (kc):
determined by the ratio between the mass balanced copper flow rate
in the concentrate and the copper flow rate in the feed: k
Rf ¼ ð5Þ
kc
mCu concentrate
Ro ¼  100 ð2Þ
mCu feed Froth recovery at each point on Fig. 3 can therefore be calculated by
dividing the measured overall flotation rate (k) at each condition by the
The overall flotation rate (k), was calculated from the overall recov- interpolated collection zone flotation rate (kc). This method enables
ery by assuming the system is a perfect mixer (Finch and Dobby, 1990), pulp effects to be decoupled from froth effects, and hence it is possible
in which k is a function of the cell residence time (τ) and the recovery to evaluate whether the tested changes either affect the pulp zone or
(R) measured in the process: froth zone responses. It should be noted that this technique, as demon-
strated by Seaman et al. (2006), estimates a froth recovery that only
R represents the loss in recovery within the froth itself and does not in-
k¼ ð3Þ clude any losses that might occur at the pulp-froth interface as the par-
τ  ð1−RÞ
ticle-laden bubbles decelerate and form the froth.
To be able to compare the overall cell performance with both froth
Both cell aspect ratio tests were operated with the same feed flow and collection zone performance simultaneously, a new method of rep-
rate (80 m3 h−1). Due to the lower cell volume, the low cell aspect resentation was adopted. This involved plotting overall flotation rate (k)
ratio resulted in a shorter residence time (1.9 min, versus 2.3 min for against froth recovery (Rf), calculated using Eq. (5). The magnitude of
the high aspect ratio). Therefore, to allow a consistent comparison, the the collection zone flotation rate (kc), by re-arranging Eq. (5), is the
overall flotation rates were first calculated using the actual residence slope of the fitted line that passes through the experimental data points
time of each test and the measured overall copper recovery (Eq. (3)). and intercepts zero (Fig. 4). Hence, the steeper the slope the higher the
The overall flotation rate was then used to recalculate the overall copper collection zone flotation rate (kc).
recovery in the low cell aspect ratio for a residence time of 2.3 min (Rτ=
2.3 min), as follows: 3. Results

kτ 3.1. Overall cell performance and reproducibility assessment


Rτ¼2:3 min ¼ ð4Þ
1þkτ
The experimental data collected generally proved to be very consis-
The determination of the collection zone flotation rate (kc) was tent and could be mass balanced relatively easily in terms of solid flow
achieved by operating the cell at three different froth depths for each rate, percent solids and copper grades.
set of combination of cell aspect ratio, impeller size, impeller tip speed In order to assess the reproducibility of the tests performed using the
and air flowrate. test rig, the rig was operated for a period of 6 h using the same operating
To determine the froth depth associated with each cell level, the conditions, first when the cell was in the high aspect configuration and
total froth depth was measured at each cell condition using a ball then again when the cell was modified to a low aspect configuration.
float. All the measured froth depths were correlated with the cell Samples in each case were taken at one-hour intervals. There was
level, and a linear model was built for each cell aspect ratio. The froth some evidence of changing operation over the six hour period of testing,
depth was then predicted based on the instrumented cell level. with variation in the copper feed grade, a decrease in the concentrate
For each froth depth, the overall flotation rate (k) was determined as solids flow rate and an increase in the copper concentrate grade. Inter-
described in Eq. (3). Thus, by applying the linear relationship between estingly, however, the copper overall recovery remained relatively con-
overall flotation rate and froth depth (Fig. 3) observed by Feteris et al. stant. It was concluded that the change in results were a consequence of
(1987) and Vera et al. (1999), the collection zone rate (kc) could be pre- a reduction in water recovery which reduced entrainment flow.
dicted by extrapolating this relationship to zero froth depth. At this This paper involves an analysis of the copper overall recovery, which
point, there should be no loss in recovery due to the froth phase and was reproducible: copper overall recovery and overall flotation rate re-
therefore the flotation rate at zero froth depth should be the collection sults had a standard deviation of around 1.3% and 0.08 min−1,
zone flotation rate. respectively.

Fig. 3. Overall flotation rate versus froth depth relationship used to estimate the collection Fig. 4. Overall flotation rate and froth recovery relationship, where “a” has higher
zone flotation rate constant (after Vera et al., 1999). collection zone flotation rate than “b”.
E. Tabosa et al. / International Journal of Mineral Processing 156 (2016) 99–107 103

Reproducibility was also assessed by running repeat tests for a num- kc ¼ 8:86−6:92AR−0:30IS þ 1:25ISz−0:036Air−1:37ðAR  ISzÞ
ber of different operating conditions (with three froth depths per condi- þ 0:043ðAR  AirÞ ð7Þ
tion) on different days of operation. Results showed a standard
deviation of 1.7% for overall copper recovery and 0.12 min−1 for overall
flotation rate values.
Finally, reproducibility was also assessed by collecting samples from Fig. 5 shows the predicted and experimental results for overall flota-
a standard condition sampled at the start of each day of testing. The tion rate (k) and collection zone flotation rate (kc). The dashed lines
standard deviation of these results was 2% for overall copper recovery show the 95% prediction interval of the model. It can be seen that the
and 0.31 min−1 for the overall flotation rate. This bigger variation for model fits the experimental results very well.
the day-to-day reproducibility tests is probably due to variation created An analysis of the error of the regression was also done. The standard
by changes in operation of the Northparkes plant (i.e. grind size, reagent error of the regression (S) is used as a measure of model fit in regression.
use). In summary, it is concluded that the test rig was capable of produc- S is measured in the units of the response variable and represents the
ing repeatable copper recovery results. standard distance data values fall from the regression line, or the stan-
Some variability in head grade between tests was expected due to dard deviation of the residuals. In this case, S was equal to 0.064 and
day-to-day changes in the feed material to the rig. Variability in head 0.2 min−1 for k and kc, respectively. This means a calculated confidence
grade on the same day was however not expected to be high because interval (2S) of 0.13 and 0.4 min−1 on the predicted k and kc results, re-
the rig was operated in recycle mode. It was therefore anticipated that spectively. The standard deviation in overall flotation rate (k) is in
the intra-day head grade would remain nearly constant. agreement with the error calculated using the reproducibility tests.
However, the head grade was found to vary between test conditions Therefore, this model can be used to eliminate the influence of the var-
within each test day. In agreement with previous studies (Runge et al., iation in copper feed grade observed during testing and to investigate
2007), an analysis of the data showed that the head grade variability be- the results of the test work.
tween tests on the same day was not random. The head grade was This was done by fixing the copper feed grade (‘Grade’) in Eq. (6) at
found to be strongly influenced by the froth depth, especially for the 0.3% Cu. This value corresponds to the average copper feed grade during
low cell aspect ratio. Deeper froths were associated with lower head the Northparkes campaign. Once the variation in copper feed grade was
grades. Copper is distributed between three regions of the test rig – eliminated, the overall flotation rate (k) and collection zone flotation
the sump, the pulp phase of the flotation cell and the froth phase of rate (kc) were predicted for different operational conditions. By doing
the flotation cell. It is believed that deeper froths result in more copper so, the overall copper recovery and froth recovery could also be predict-
being contained within the froth phase of the flotation cell thereby caus- ed using Eqs. (4) and (5) with cell residence time (τ) equal to 2.3 min.
ing a drop in the average grade of the feed in the sump (since the total The standard deviations of the calculated overall copper recovery
amount of copper in the system must remain constant). and froth recovery were estimated using a Monte Carlo technique.
Variability in head grade would make analysis of the metallurgical re- This procedure is a practical and efficient mean of estimating error prop-
sults problematic as it may have a significant impact on flotation perfor- agated from input parameters of known standard deviations (Ogilvie,
mance. Therefore, in order to compare the metallurgical performance 1984). It randomly creates a normally distributed dataset based on the
with no variation in copper feed grade, a model was developed for overall input's mean and standard deviation.
flotation rate (k) which incorporated a copper feed grade parameter (see The technique was used to create 1000 random values of k and kc.
below). Feed grade could then be kept constant in subsequent analysis. Overall copper recovery and froth recovery were then predicted and
their standard deviations calculated. It is estimated that overall copper
3.2. Overall flotation rate and collection zone flotation rate modelling recovery would have a standard deviation of 2.1%, which is in agree-
ment with the standard deviation found from the analysis of the repro-
In order to compare the metallurgical performance with no variation ducibility tests. Froth recovery was calculated to have a standard
in copper feed grade, a model was developed for overall flotation rate deviation of 10.6%. The standard deviation of froth recovery is expected
(k) and collection zone flotation rate (kc). Overall and froth recoveries to be higher than that of overall copper recovery. This is because froth
can then be calculated using Eqs. (4) and (5) that are directly compara- recovery has been calculated based on an extrapolation method that
ble, allowing evaluation of whether the variables changed in the test will be sensitive to any error in the froth depth and overall flotation
program affect the pulp or froth zone responses or both. rate measurements.
Multiple regression was used to build the models. Stepwise linear
regression was chosen, as it generates an optimal equation for 3.3. Effect of cell hydrodynamic on flotation performance
predicting a dependent variable from several independent variables.
The dependent variables were the overall flotation rate (k) and collec- Plotting overall flotation rate against froth depth (as described for
tion zone flotation rate (kc). The independent variables were the cell as- Fig. 4) allows the interactions between overall cell performance and
pect ratio (AR), impeller tip speed (IS), impeller mechanism size (ISz), both froth and collection zone performance to be evaluated
air flow rate (Air), froth depth (FD) and copper feed grade (Grade). simultaneously.
The systematic iterative construction of the regression model in-
volved the selection of the independent variables, and their interactions, 3.3.1. Effect of impeller size
by including all variables in the model and eliminating those that are Fig. 6 shows that for every cell aspect ratio, impeller tip speed com-
not statistically significant (95% confidence interval). Minitab statistical bination, the collection zone flotation rate of the oversized RCS5 impel-
software was used to optimise the model. By successively adding or re- ler mechanisms was higher than the standard RCS3. This difference was
moving variables based on the t-test statistics of their estimated coeffi- more significant for the low cell aspect ratio (around 30% and 8% higher
cients the linear model for overall flotation rate (k) and collection zone for low and high cell aspect ratio respectively). It is anticipated that, for
flotation rate (kc) were found (Eqs. (6) and (7)). the same impeller tip speed, the oversized impeller mechanism will dis-
k ¼ −5:66 þ 13:75AR þ 7:64Grade−0:18ðAR  ISÞ−0:58ðAR  ISzÞ sipate more energy than the standard impeller mechanism. Therefore, it
will result in higher local energy dissipation rates in the impeller stream
−0:033ðAR  AirÞ−0:16ðAR  FDÞ−11:58ðAR  GradeÞ and thus higher particle-bubble collision rates (Schubert, 2008).
ð6Þ
þ 0:018ðIS  ISzÞ−0:24ðIS  GradeÞ þ 0:0019ðISz  AirÞ However, froth recovery tends to be lower for the oversized impeller
(RCS5) than the standard RCS3 mechanism and this is presumably due
þ 0:00093ðAir  FDÞ þ 0:037ðAir  GradeÞ to instability of the froth base caused by greater turbulence near the
104 E. Tabosa et al. / International Journal of Mineral Processing 156 (2016) 99–107

Fig. 5. Predicted versus experimental overall flotation rate (k) and collection zone flotation rate (kc).

froth pulp interface. This resulted in lower overall recoveries for the There is an interactive effect between air rate and aspect ratio that is
oversized RCS5 impeller. significant. At the low cell aspect ratio, high air flowrate resulted in a
slightly higher overall flotation rate and froth recovery.
3.3.2. Effect of impeller tip speed
Froth recovery was not overly affected by impeller tip speed. For low 3.4. Correlating cell hydrodynamics with collection zone flotation
and high aspect ratio, the collection zone flotation rate at high impeller performance
tip speed was lower than at low impeller tip speed (Fig. 7).
The reduction in collection zone flotation rate with an increase in Many authors have reported that flotation performance is propor-
impeller speed is not consistent with the increase in collection zone flo- tional to the bubble surface area flux (Sb), notably in the k-Sb relation-
tation rate constant observed with an increase in impeller size. Both an ship proposed by Gorain et al. (1998). In the present study, however,
increase in impeller tip speed and an increase in impeller size increase it was demonstrated that the different cell hydrodynamic conditions
the amount of energy dissipated in the flotation cell and it would there- do affect flotation performance in the collection zone.
fore be expected that the direction of change in performance would be It seems therefore that for the k-Sb relationship to be valid in this
the same (not the opposite). These results are an indication that it is not particular study, some term related to the hydrodynamic conditions in
just total energy input that affects flotation response. the cell should be included as kc = P ∙ f(hydrodynamics) ∙ Sb, where P is
the floatability of the particles which is a function of the ore and chem-
ical conditions in the cell (Gorain, 1997 and Gorain et al., 1998). A sim-
3.3.3. Effect of cell aspect ratio ilar approach has been proposed by Amini et al. (2015).
Fig. 8 shows that for low and high impeller tip speeds, the collection By comparing the collection zone flotation rate (kc) with cell overall
zone flotation rate of the low aspect ratio cell was greater than the high energy dissipation (Fig. 10a), it can be seen that kc was not related to
aspect ratio cell, presumably due to the greater proportion of turbulence overall energy input alone as is commonly observed at laboratory
within the cell. scale, where energy is usually only changed by varying impeller speed.
The higher turbulence closer to the froth pulp interface adversely af- Correlations, however, were observed between collection zone flotation
fects froth recovery, which is about 40 to 50% lower in the low cell as- rate and overall energy dissipation, for each cell aspect ratio and impel-
pect ratio than in the high cell aspect ratio. This is presumably due to ler mechanism size combination (Fig. 10b).
instability of the froth base. This may suggest that flotation rate in the collection zone (kc) is pro-
portional not only to the amount of energy being dissipated by the im-
3.3.4. Effect of air flowrate peller but also to the way energy is dissipated. For example, an increase
For a particular cell aspect ratio and impeller tip speed, the collection in energy dissipation by increasing the ratio between high and low tur-
zone flotation rate (kc) and froth recoveries were similar for both low bulent zones (i.e., lower cell aspect ratio or oversized impeller mecha-
and high air flowrates (Fig. 9). nism) significantly increases the collection zone flotation rate. Whilst

Fig. 6. Effect of impeller size on the flotation performance.


E. Tabosa et al. / International Journal of Mineral Processing 156 (2016) 99–107 105

Fig. 7. Effect of impeller tip speed on the flotation performance.

an increase in energy dissipation by increasing impeller tip speed tends conditions is 10% higher than the standard conditions under which
to reduce the collection zone flotation rate. the flotation cell is usually operated (i.e. high aspect ratio and high im-
Therefore, better flotation performance in the collection zone might peller speed).
be achieved when, for a given total energy dissipation rate, more of the
energy is dissipated in the impeller-stator region (e.g. by using a low cell 4. Discussion
aspect ratio or increasing impeller mechanism size). Improvements to
the froth phase performance are also necessary to avoid losses in this It has been demonstrated that cell hydrodynamics play an important
zone. role in flotation performance. The results presented in this paper have
To make a quantitative assessment of how an improved collection the following implications for flotation cell design, scale-up and
zone flotation rate would translate into an overall recovery improve- operation:
ment without the froth recovery effect, Eq. (9) was used to calculate
the collection zone recovery (Rc). The residence time used in Eq. (9) • An increase in the size of the high turbulent zone results in an increase
was 2.5 and 1.9 min for the high and low cell aspect ratio, respectively. in the rate of flotation in the collection zone of a flotation cell. The pro-
The curves that resulted from this analysis are shown (dotted lines) in portion of the cell that is highly turbulent has decreased significantly
Fig. 11. Overall recovery (Ro) in a flotation cell is a function of recovery as the size of flotation cells has dramatically increased in size over
in the collection (Rc) and froth (Rf) phase, (Dobby, 1984): the last 20 years. It is likely that this has resulted in a quiescent zone
that is too large in size for what is required to protect the froth zone
R f Rc from the turbulence of the impeller. The height of the cell could be re-
Ro ¼ ð8Þ duced without an adverse effect on the overall recovery achieved in
R f Rc þ 1−Rc
the flotation cell – thus reducing the amount of steel used and there-
fore the cost of manufacture of the cell. Alternatively, in the larger flo-
kc τ tation machines multiple impellers (possibly with different designs)
Rc ¼ ð9Þ
1 þ kc τ at different heights would potentially result in a greater proportion
of the cell being highly turbulent, resulting in higher flotation recover-
The collection zone copper recovery (Rc) may therefore be obtained ies.
when there is no froth recovery effect, i.e. when froth recovery equals • To be able to maximise the turbulence used in a flotation cell, design
100%. By comparing the collection zone copper recoveries for a particu- features which minimise or protect the flotation froth from the turbu-
lar cell aspect ratio, it can be seen that the low impeller tip speed in- lence effects should be implemented. For example internal baffle rings
creased the collection zone recovery by 3% for the low cell aspect ratio around the cell circumference located above the impeller region can
and 6.4% for the high cell aspect ratio. be used to divert flow so it does not travel upwards and disrupt the
Best overall performance would have been achieved with a low cell froth zone. Alternatively, some sort of turbulence diffuser could be
aspect ratio running at a low impeller speed. Recovery under these placed below the froth to minimise turbulence effects.

Fig. 8. Effect of cell aspect ratio on the flotation performance.


106 E. Tabosa et al. / International Journal of Mineral Processing 156 (2016) 99–107

Fig. 9. Effect of air flowrate on the flotation performance (RCS3 impeller).

• Oversized mechanisms can potentially result in increased collection turbulence distribution to be measured as a function of cell design and
zone rates and are therefore an option worth evaluating. operating conditions, and for the results to be correlated with flotation
• In terms of cell operation, it is likely that the optimal impeller speed performance. Further work will be required to elucidate the mecha-
will be different, depending on the particle size and hydrophobicity nisms by which impeller speed and size, and cell aspect ratio affect col-
of the ore being processed. Mechanical flotation cells should be pro- lection and froth phase recovery rates.
vided with a variable speed drive motor in order to optimise the im-
peller speed.
• Flotation recovery rates would be expected to be higher in a smaller 5. Conclusions
volume cell than a larger volume cell (same aspect ratio) because a
larger proportion of the cell is highly turbulent. The aim of this study was to investigate the effects of cell hydrody-
• Cell hydrodynamic conditions could be optimised in order to improve namics on flotation kinetics in a pilot-scale flotation cell. The metallur-
coarse (less liberated) particle recoveries. Evidence suggests that gical experimental data collected proved to be very consistent and,
coarse particle recovery is extremely sensitive to froth phase effects besides the variability with regards to the copper feed grade observed
with recovery optimal at shallow froth depth and when turbulence throughout the campaign, tests showed that the test rig was capable
at the pulp-froth interface is minimised (Tabosa et al., 2013). Further of producing very reproducible results.
study on the effect of cell hydrodynamics on flotation performance of Although the overall copper recoveries were very similar for the
coarse and less liberated particles is recommended. conditions tested, by decoupling the pulp effect from the froth effect it
was possible to evaluate whether the changes tested affected the collec-
Understanding how cell hydrodynamic conditions affect perfor- tion zone response, the froth zone response or both. The analysis
mance potentially allows them to be better controlled, thereby provid- showed that the overall recovery had the potential to be up to 10%
ing an additional means by which flotation can be enhanced. higher if not limited by froth recovery.
Furthermore, because turbulent conditions inside a flotation cell have The cell aspect ratio was the variable that had the most notable effect
usually only been changed by varying impeller tip speed and/or air on copper collection zone recovery. A change from the standard high
flow rate, there remains a need to better understand the relationship cell aspect to the low cell aspect ratio, however, had little effect on the
between changes in impeller tip speed combined with changes in im- overall cell performance. A detailed analysis showed that the high cell
peller/stator mechanism size and cell design. The results in this paper aspect ratio had a lower collection zone recovery, but was operating at
suggest that it is the turbulence distribution within the flotation cell, higher froth recoveries. Whereas the low cell aspect ratio presented
rather than just the energy input, that needs to be optimised. higher collection zone recovery, but its overall recovery was limited
New techniques have been developed to measure the distribution of by the low froth recovery.
turbulence in an abrasive opaque slurry inside a flotation cell (Tabosa, The effect of other operational variable tested can be summarised as
2012; Tabosa et al., 2012; Meng et al., 2015). These will allow the follows:

Fig. 10. Effect of energy dissipation on collection zone flotation rate for (a) ungrouped data and (b) grouped by aspect ratio and impeller size.
E. Tabosa et al. / International Journal of Mineral Processing 156 (2016) 99–107 107

Feteris, S.M., Frew, J.A., Jowett, A., 1987. Modelling the effect of froth depth in flotation.
Int. J. Miner. Process. 20, 121–135.
Finch, J.A., Dobby, G.S., 1990. Column Flotation. Firth ed. Pergamon Press, Oxford.
Gorain, B.K., 1997. The Effect of Bubble Size on the Kinetics of Flotation and Its Relevance
to Scale-up PhD Thesis University of Queensland, Julius Kruttschnitt Mineral Research
Centre, Australia.
Gorain, B.K., Napier-Munn, T.J., Franzidis, J.P., Manlapig, E.V., 1998. Studies on impeller
type, impeller speed and air flow rate in an industrial scale flotation cell. Part 5: val-
idation of k-Sb relationship and effect of froth depth. Miner. Eng. 11 (7), 615–626.
Hall, G.A., Dell, C.C., 1989. An experimental study of the role of turbulence in flotation cell
design. Mineral Processing in the United Kingdom. Proceedings of a Meeting Orga-
nized by the Institution of Mining and Metallurgy. 4–5 April, 1989, pp. 149–164.
Jameson, G.J., Ahmed, N., 1983. Improving the rate of flotation of fine particles. Chemeca
83: The Eleventh Australian Chemical Engineering Conference. Brisbane, Australia,
pp. 203–208.
Jordan, C.E., Spears, D.R., 1990. Evaluation of a turbulent flow model for fine-bubble and
fine particle flotation. Miner. Metall. Process. 65–73 May.
Koh, P.T.L., Manickam, M., Schwarz, M.P., 2000. CDF simulation of bubble-particle colli-
sions in mineral flotation cells. Miner. Eng. 13 (14–15), 1455–1463.
Meng, J., Xie, W., Tabosa, E., Runge, K., Bradshaw, D., 2015. Turbulence modelling for flo-
Fig. 11. Overall copper recovery versus froth recovery relationship. tation cells based on piezoelectric sensor measurement data. Proceedings of the 7th
International Flotation Conference (Flotation '15), Cape Town, South Africa.
Morris, R.M., Matthesius, G.A., 1988. Froth flotation of coal fines - the influence of turbu-
• Air flow rate (for the range evaluated here) had little effect on the flo- lence on cell performance. J. South. Afr. Inst. Min. Metall. 88 (12), 385–391.
tation performance; Newell, R., Grano, S., 2006. Hydrodynamics and scale up in Rushton turbine flotation cells:
• Low impeller tip speed had better flotation performance than the high part 2. Flotation scale-up for laboratory and pilot cells. Int. J. Miner. Process. 81 (2),
65–78.
impeller tip speed (within the range used in this study); Newell, R., Grano, S., 2007. Hydrodynamics and scale up in Rushton turbine flotation cells:
• The oversized impeller mechanisms (RCS5) resulted in higher collec- part 1 - cell hydrodynamics. Int. J. Miner. Process. 81 (4), 224–236.
tion zone flotation rate compared to the standard RCS3 mechanism, Ogilvie, J.F., 1984. A Monte-Carlo approach to error propagation. Comput. Chem. 8 (3),
205–207.
when operating at the same impeller tip speed and cell aspect ratio. Pyke, B.L., Duan, J., Fornasiero, D., Ralston, J., 2002. From turbulence and collision to at-
The RCS5, however, resulted in poorer flotation performance when tachment and detachment: a general flotation model. In: Ralston, J., Miller, J.D.,
compared to the RCS3 because the overall flotation rate and froth re- Rubio, J. (Eds.), Flotation + Flocculation: From Fundamentals to Applications,
pp. 77–89 Hawaii.
covery decreased.
Pyke, B., Fornasiero, D., Ralston, J., 2003. Bubble particle heterocoagulation under turbu-
lent conditions. J. Colloid Interface Sci. 265, 141–151.
Comparing the metallurgical performance of the cell with the differ- Runge, K.C., Seaman, B., Seaman, D., 2007. Optimising down the bank flotation froth per-
ent hydrodynamic conditions it was found that the collection zone flo- formance in Metso's 3 m3 RCS™ test cell. Flotation '07. Minerals Engineering Confer-
ence, Cape Town, South Africa.
tation rate was not directly related to overall energy dissipation, as is Rutherford, K., Mahmoudi, S.M.S., Lee, K.C., Yianneskis, M., 1996. The influence of rushton
commonly observed at laboratory scale, where energy is usually only impeller blade and disk thickness on the mixing characteristics of stirred vessels.
changed by varying impeller tip speed. Results suggest that it is the Chem. Eng. Res. Des. 74 (3), 369–378.
Schubert, H., 1985. On some aspects of the hydrodynamics of flotation processes. In:
size of the turbulent zone, rather than just energy input, that affects flo- Forssberg, E. (Ed.), Flotation of Sulphide Minerals, pp. 337–352.
tation recovery. Schubert, H., 1986. On the hydrodynamics and the scale-up of flotation processes. In:
Somasundaran, P. (Ed.), Advances in Mineral Processing. Soc. Min. Eng., Littleton
(636–349).
Acknowledgements Schubert, H., 1999. On the turbulence-controlled microprocesses in flotation machines.
Int. J. Miner. Process. 56, 257–276.
Schubert, H., 2008. On the optimization of hydrodynamics in fine particle flotation. Miner.
Professor Dr. Heinrich Schubert and his team are gratefully acknowl- Eng. 21 (12–14), 930–936.
edged for laying the foundation for the understanding of the role of cell Schubert, H., Bischofberger, C., 1978. On the hydrodynamics of flotation machines. Int.
hydrodynamics on flotation performance. The authors would also like J. Miner. Process. 5 (2), 131–142.
Schubert, H., Bischofberger, C., 1979. On the optimization of hydrodynamics in flotation
to acknowledge Metso Process Technology and Innovation for funding
processes. XIII International Mineral Processing Congress, Part B, pp. 353–375.
this research and the assistance of personnel involved in the testing Schubert, H., Bischofberger, C., 1998. On the microprocesses air dispersion and particle-
campaign and during sample preparation and data analysis: Jaclyn bubble attachment in flotation machines as well as consequences for the scale-up
of macroprocesses. Int. J. Miner. Process. 52 (4), 245–259.
McMaster, Robert Crosbie, Kevin Cummins and Rae de Rusett. Rio
Schubert, H., Bischofberger, C., Koch, P., 1982. On the influence of the hydrodynamics in
Tinto's Northparkes operation is also gratefully acknowledged for flotation processes. XIV International Mineral Processing Congress. Toronto, Canada,
assisting during the research work and assay analysis. pp. IV-15.1–IV-15.17.
Seaman, D.R., Manlapig, E.V., Franzidis, J.P., 2006. Selective transport of attached particles
across the pulp–froth interface. Miner. Eng. 19 (6–8), 841–851.
References Spears, D.R., Jordan, C.E., 1989. The effect of turbulence on the flotation of galena when
using fine bubbles. Advances in Coal and Mineral Processing Using Flotation, pp. 77–84.
Ahmed, N., Jameson, G.J., 1985. The effect of bubble size on the rate of flotation of fine par- Tabosa, E., 2012. The Effect of Cell Hydrodynamics on Flotation Kinetics PhD Thesis Julius
ticles. Int. J. Miner. Process. 14 (3), 195–215. Kruttschnitt Mineral Research Centre, The University of Queensland, Brisbane,
Amini, E., Xie, W., Bradshaw, D., 2015. Enhancement of the precision and accuracy of the Australia.
flotation scale up process by introducing turbulence parameters to the AMIRA P9 Tabosa, E., Runge, K., Holtham, P., 2012. Development and application of a technique for
model. Proceedings of the 7th International Flotation Conference (Flotation '15), evaluating turbulence in a flotation cell. Proceedings of the XXVI International Mineral
Cape Town, South Africa. Processing Congress, IMPC 2012, New Delhi, India, 24–28 September, pp. 5377–5390.
Cheng, Y.H., Mikhail, M.W., Salama, A.I.A., 1995. Effect of turbulence on the flotation of Tabosa, E., Runge, K., Duffy, K., 2013. Strategies for increasing coarse particle flotation in
fine coals. In: Laskowski, J.S.P., G.W. (Eds.), 1st UBC-McGill Bi-annual International conventional flotation cells. Flotation '13. Minerals Engineering Conference, Cape
Symposium on Fundamentals of Mineral Processing, Vancouver, pp. 425–434. Town, South Africa.
Deglon, D.A., Sawyerr, F., O'Connor, C.T., 1999. A model to relate the flotation rate constant Vera, M.A., Franzidis, J.P., Manlapig, E.V., 1999. Simultaneous determination of collection
and the bubble surface area flux in mechanical flotation cells. Miner. Eng. 12 (6), zone rate constant and froth zone recovery in a mechanical flotation environment.
599–608. Miner. Eng. 12, 1163–1176.
Deglon, D.A., Egya-mensah, D., Franzidis, J.P., 2000. Review of hydrodynamics and gas dis- Weber, A., Walker, C., Redden, L., Lelinski, D.S., Ware, S., 1999. Scale-up and design of large
persion in flotation cells on South African platinum concentrators. Miner. Eng. 13 (3), scale flotation equipment. In: Parekh, B.K., Miller, J.D. (Eds.), Advances in Flotation
235–244. Technology. SME, Littleton, CO, pp. 353–370.
Dobby, G.S., 1984. A fundamental flotation model and flotation column scale-up. Depart- Weiss, T., Schubert, H., 1988. The effects of fine particles on the hydrodynamics of flota-
ment of Mining and Metallurgical Engineering. McGill University, Montreal, p. 259 tion processes. In: Forssberg, E. (Ed.), XVI International Mineral Processing Congress,
PhD thesis. pp. 807–818.
Duan, J., Fornasiero, D., Ralston, J., 2003. Calculation of the flotation rate constant of chal- Wu, H., Patterson, G.K., 1989. Laser-Doppler measurements of turbulent-flow parameters
copyrite particles in an ore. Int. J. Miner. Process. 72 (1–4), 227–237. in a stirred mixer. Chem. Eng. Sci. 44 (10), 2207–2221.

View publication stats

You might also like