Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Int. J. Miner. Process.

72 (2003) 255 – 266


www.elsevier.com/locate/ijminpro

A study of bubble coalescence in flotation froths


Seher Ata *, Nafis Ahmed, Graeme J. Jameson
Centre for Multiphase Processes, University of Newcastle, University Drive, Callaghan, Newcastle, NSW 2308, Australia

Received 19 January 2003; received in revised form 19 June 2003; accepted 1 July 2003

Abstract

This paper is concerned with changes in bubble size and bubble size distribution in the froth phase of a flotation column. A
continuous flotation cell of special design is used in which deep froths can be formed. The effect of parameters such as degree
of hydrophobicity, gangue concentration (entrained solids), initial bubble size (pulp bubble size) and froth height has been
investigated. Special attention has been given to the use of particles of well-defined hydrophobicities so that their effect on the
behaviour of the froth phase could be assessed more accurately. Glass particles of different hydrophobicities (as reflected in the
contact angle) are prepared using controlled silanation. Contact angles chosen are 50j, 66j and 82j. The results suggest that the
size of the bubbles strongly depends on the degree of hydrophobicity of particles. In the presence of entrained solids, the bubble
coalescence rate substantially decreases mainly due to reduced liquid drainage in the bubble films.
D 2003 Elsevier B.V. All rights reserved.

Keywords: froth flotation; bubble size; froth stability; froth properties

1. Introduction However, a flotation froth can exhibit a quite different


structure from a two-phase foam since it contains
Despite the importance of froth in the flotation particles that in turn may play a key role in bubble
process, little attention has been paid to the details of coalescence and froth stability. The behaviour of the
the mechanisms, which governs the bubble behaviour flotation froth therefore is governed not only by the
in the froth column. The process is very difficult to factors that rule two-phase foams such as film thinning
study because the presence of particles in the froth or gravity drainage but also by contributions from the
column prevents the measurement of system parame- particles. The study of each process is equally impor-
ters, such as gas holdup or even bubble size, accurate- tant in understanding the mineralized froth.
ly. As a result, attempts to model the flotation froth at a The aim of the work reported herein is to obtain a
macroscopic level rely much more on the knowledge better understanding of froth behaviour under differ-
of two-phase froths rather than actual phenomena. ent conditions, through observation of the changes of
the bubble size distribution in the rising froth due to
coalescence. Special attention has been given to
* Corresponding author. Tel.: +61-2-4921-6181; fax: +61-2-
particle hydrophobicity. The parameters of interest
4960-1445. are the hydrophobicity of the solid particles, the initial
E-mail address: Seher.Ata@newcastle.edu.au (S. Ata). bubble size and the concentration of gangue mineral.

0301-7516/$ - see front matter D 2003 Elsevier B.V. All rights reserved.
doi:10.1016/S0301-7516(03)00103-0
256 S. Ata et al. / Int. J. Miner. Process. 72 (2003) 255–266

2. Experimental ing, Australia. The average particle size (d50) was 68


Am; the d80 was 82 Am.
2.1. Materials Glass particles of different hydrophobicities (as
reflected in the contact angle) were prepared using
2.1.1. Silica controlled silanation. Contact angles (h) chosen were
Washed quartz obtained from Commercial Miner- 50j, 66j and 82j. Details of the preparation of glass
als, Australia, under the brand name Silica #400 was particles may be found in Ata et al. (2002).
used as the hydrophilic (gangue) solid. The average
particle size of the sample (d50), determined using a 2.2. The flotation apparatus
Malvern Particle Sizer, was 5 Am.
A schematic of the cell is shown in Fig. 1. The
2.1.2. Glass spheres flotation cell consists of a 2.3-l, 150-mm diameter
Glass spheres were used as the hydrophobic model stirred vessel connected to a 50-mm diameter column
system. The glass particles used in this study were through a tapered transition piece. The transition piece
Mill Spec-13 grade supplied by Godfrey Blast Clean- was necessary in order to achieve a formation of deep

Fig. 1. Flotation cell used to measure bubble size in the froth.


S. Ata et al. / Int. J. Miner. Process. 72 (2003) 255–266 257

froth over wide operating conditions. The collection (d32) is defined as the volume-to-surface mean bubble
zone was constructed from Perspex, and the column diameter:
section was made of glass to which bubbles do not P 3
adhere. Bubbles were generated by introducing nitro- ni d
d32 ¼ P i2 ð1Þ
gen through a sintered glass frit incorporated into the ni di
base of the cell. The bubble size was varied by using
frits with different pore sizes. where ni is the number of bubbles with diameter di.
The flotation cell was operated in a continuous The Sauter-mean bubble diameter is commonly used
mode. A batch of feed was stirred in a 25-l vessel for in flotation studies. This usage arises because the rate
8 min and then fed into the cell through a constant head of flotation is closely linked to the surface area of the
tank. A constant flow rate of slurry through the cell bubbles.
was achieved by the use of a peristaltic pump, which The axial profiles of the Sauter-mean diameter
was set to deliver the required flow of 650 ml/min. The (d32) through the bubble bed in the presence of solids
tailing stream was passed through a gravity overflow. of different hydrophobicities are shown in Fig. 2.
The gravity overflow was also used to maintain the Three levels of hydrophobicity were investigated,
position of the froth liquid interface at a preset level in i.e., strongly, intermediate and weakly corresponding
the column. The superficial gas velocity was kept to contact angles of 82j, 66j and 50j, respectively.
constant at 1.2 cm/s, and frother Dowfroth 250, at a Tests were conducted at 1.2 cm/s superficial gas
constant dosage of 30 mg/l, was the only reagent added velocity; 320 mm of froth depth; 30 ppm of frother
to the cell during the experiments. concentration; 650 cm3/min feed flow rate; and 7.5%
In the present study, the bubble size was deter- solids concentration. It should be noted that, except
mined by a photographic technique. Bubbles near the for the parameter under study, these experimental
column wall in the froth zone were photographed conditions were kept unchanged in all of the tests.
using a Nikon camera after the concentrate reached a However, under the various operating conditions, it
steady state. Distortion of images due to the cylin- was observed that the froths formed above the pulp
drical surface of the column was eliminated by zone were generally different in height. Thus, for each
employing a transparent viewing box filled with system studied, the froth depth was adjusted to allow a
water around the column. The box could be moved free flow of froth over the cell lip.
up and down easily to allow bubbles to be photo- It is evident from Fig. 2 that the Sauter-mean bubble
graphed at various levels in the froth. A graduated diameter, in the presence of hydrophobic particles,
scale was secured to the front of the box to provide increases with increasing distance from the pulp – froth
the necessary scale. Tests indicated that side lighting interface. The major increase in d32 occurs with
by the flash gave the best definition. The flash was particles of the lowest contact angle; these particles
covered by a triple layer of tracing paper. The appear to have less effect on the stability of the froth.
camera setting was f/8 and 1/100 s shutter speed. Interestingly, the particles that have greatest bubble
Black-and-white Agfa 25 ISO films were used to stability were those of intermediate hydrophobicity.
provide extra definition. At least 500 bubbles were In order to observe the effect of entrained particles
sized in each case. The negatives of the photographs on the bubble size in the froth phase, a series of tests
were projected using an enlarger directly onto a were conducted with feeds containing a mixture of
digitiser pad (Summagraphics) and sized using a hydrophobic glass spheres and gangue mineral (sili-
personal computer. ca). The results are shown in Fig. 3 where the Sauter-
mean bubble diameter is plotted as a function of the
height above the pulp – froth interface for particles of
3. Results the three hydrophobicities. The total concentration of
solids in the feed was 15% (w/w), with equal masses
In this study, the Sauter-mean diameter is used as of glass and silica; the froth depth was at 450 mm.
the representative size for the bubble size distribution The remaining operating conditions are as given for
in the froth phase. The Sauter-mean bubble diameter Fig. 2. Silica did not exhibit any floatability, so it was
258 S. Ata et al. / Int. J. Miner. Process. 72 (2003) 255–266

Fig. 2. Variation of the Sauter-mean bubble diameter as a function of froth height for glass having various degree of hydrophobicities with the
Sparger ‘1’ bubbles.

transported into the froth by entrainment. Bubble size of silica in the flotation cell for these tests was 7.5%
measurement was also carried out in the system where (w/w). Results are given in Fig. 3.
only hydrophilic particles (silica) were present to It is seen that the d32 values obtained with the
determine whether the hydrophobic glass particles mixed system show a trend similar to that observed
do in fact play a role in all cases. The concentration with the system in which only hydrophobic spheres

Fig. 3. The Sauter-mean bubble diameter in the froth for glass of different hydrophobicities and hydrophilic solid (silica) in the mixed system
with the Sparger ‘1’ bubbles.
S. Ata et al. / Int. J. Miner. Process. 72 (2003) 255–266 259

Table 1 corresponding values in the mixed system are 1.1, 0.9


Statistical analysis of bubble size in the mixed and glass spheres and 1.3 mm, respectively.
systems for Sparger ‘1’ bubbles
Table 1 gives a statistical analysis of bubble size at
Contact angle (h) P value
the interface in the two-solid (glass spheres and silica)
h = 82j 0.0026 and the one-solid (glass spheres) systems for the three
h = 66j 0.0203
hydrophobic particles, using the F-test. Probability
h = 50j 0.2580
( P) values more than or equal to 0.05 were indicative
of the size of the bubbles in two systems being
are present. The d32 increases continuously with froth statistically different at a confidence level z 95%. It
height for all three hydrophobicities and hydrophilic is seen that with strongly and intermediate hydropho-
particles. Again, it is significant to observe that the bic glass, the bubble size at the interface with glass
increase in d32 is the lowest for intermediate hydro- spheres and mixed systems is the same at the 95%
phobic particles, while being the highest for weakly confidence level, while with weakly hydrophobic
hydrophobic particles and hydrophilic silica. Note that particles, the size of the bubbles at the pulp – froth
the bubble growth rate in the presence of hydrophilic interface is different in the two systems.
particles and weakly hydrophobic glass seems to be Tests were also carried out to investigate whether
similar, suggesting that the contribution of weakly the bubble size in the froth zone depends on the initial
hydrophobic glass to the stability of the froth is bubble size (the bubble size in the liquid phase). In the
minimal. present study, the initial bubble size was varied
A comparison between Figs. 2 and 3 shows that the independently using another sparger having a differ-
change in the Sauter-mean bubble diameter with ent porosity. For convenience, this will be designated
respect to the froth height is considerably reduced in as Sparger ‘0’. Similarly, the sparger used in Figs. 2
the mixed system for all three hydrophobicities. For and 3 will be nominated as Sparger ‘1’. Sparger ‘0’
example, in a system of hydrophobic glass particles has a larger pore size distribution than Sparger ‘1’,
alone, at a froth level of 200 mm, the d32 for the which means that under the same experimental con-
strongly, intermediate and weakly hydrophobic par- ditions, it produces larger bubbles. The comparison of
ticles is 1.6, 1.5 and 1.9 mm, respectively, while the bubble sizes produced by the two spargers was made

Fig. 4. Variation of the Sauter-mean bubble diameter as a function of froth height, for glass particles of various of hydrophobicity and
hydrophilic solid (silica) in the mixed system, with Sparger ‘0’ bubbles.
260 S. Ata et al. / Int. J. Miner. Process. 72 (2003) 255–266

Fig. 5. Bubble size distributions as a function of froth height for three hydrophobic glass particles with Sparger ‘1’ bubbles. Froth height was
320 mm and flotation feed contained only glass spheres. (a) Weakly hydrophobic glass (h = 50j); (b) Intermediate hydrophobic glass (h = 66j);
(c) Strongly hydrophobic glass (h = 82j).
S. Ata et al. / Int. J. Miner. Process. 72 (2003) 255–266 261

using the F-test in the system where only silica was


present, to eliminate the effect of particle hydropho-
bicity on the bubble size. The results indicated that at
a confidence level of 95%, the two spargers indeed
produce different bubble size distributions. It should
be mentioned that with the Sparger ‘0’ bubbles,
insufficient froth formation was obtained when only
hydrophobic glass particles were present in the feed.
This indicates that the hydrophilic particles have a
stabilizing effect on the froth, possibly by increasing
the effective viscosity, or by mechanical blockage.
(By mechanical blockage, we mean the obstructions
to the flow in the liquid films in the froth, caused by
the presence of particles held on the surfaces of the
bubbles and protruding into the liquid. This effect
would be expected to become significant when the
thickness of the films becomes of the same order as
the size of the particles.) Thus, with Sparger ‘0’,
experiments were conducted only with the mixed feed
(glass spheres + silica). Results from these tests are
shown in Fig. 4, where the Sauter-mean bubble
diameter is plotted against the height above the froth
interface. As with Sparger ‘1’ bubbles, there is sub-
stantial bubble coalescence with increasing froth
height in the presence of three hydrophobic particles
and hydrophilic solid.
Comparing Figs. 3 and 4, it appears that with the
Sparger ‘0’ bubbles, the Sauter-mean bubble diameter
obtained at various locations in the froth for the three
levels of hydrophobicities and hydrophilic particles is
higher than those obtained with the Sparger ‘1’
bubbles. It is important to observe that the Sauter-
mean diameter at the various froth heights in the
presence of hydrophilic particles is very similar to
those in the presence of weakly hydrophobic glass.
Fig. 5a– c shows typical histograms of the bubble
size at various levels in the froth zone for the weakly,
intermediate and strongly hydrophobic glass, respec-
tively. For the sake of brevity, only the results for the
froth containing hydrophobic glass spheres are shown,
but the same trends were followed in a mixed system
with both Sparger ‘0’ and Sparger ‘1’ bubbles. It is

Fig. 6. Photographs of gas bubbles in the froth in the presence of


strongly hydrophobic glass (h = 82j) with Sparger ‘1’ bubbles.
Froth height was 320 mm and flotation feed contained only glass
spheres. (a) Level, 200 mm (interface); (b) level, 100 mm; (c) level,
0 mm.
262 S. Ata et al. / Int. J. Miner. Process. 72 (2003) 255–266

seen from the figures that the Sauter-mean bubble film breakers if they have a contact angle of more
diameter increases for each with increasing distance than 80j. With particles of intermediate hydropho-
from the pulp – froth interface for all the hydrophobic bicity (corresponding to h = 65j), the froth stability
particles. The distribution at the interface has a very was maximized suggesting that there is an optimum
narrow size range. As the froth height above the particle hydrophobicity that promotes froth stability.
interface increases, the bubble size distribution From observations on thin films containing particles,
becomes wider, suggesting that large bubbles emerge, it was suggested that particles of intermediate hydro-
but also that a significant number of small bubbles phobicity were capable of forming stable bridges
remain. This may be seen clearly in Fig. 6a –c where across the film, increasing the rigidity of the froth
bubbles were photographed at different levels in the structure. Particles of high hydrophobicity (h = 82j)
froth. The interesting aspect of these pictures lies in penetrated the interface to a much greater extent and
the fact that the structure of the froth is very different ruptured the film, thus leading to unstable film
from the froth structure that is generally used in bridging. Particles with low degree of hydrophobicity
flotation froth modeling. Even at the highest level of corresponding to a contact angle of h < 40j were
the froth (Fig. 6a) where the liquid content is the found to stream out into the lamella and did not
lowest, the shape of the plateau border is not defined contribute to the stability of the froth film. Aveyard et
clearly. al. (1994) studied the effect of spherical glass beads
In Fig. 5a – c, a comparison between the bubble of varying hydrophobicity on foam stability, and
size distributions at three levels of froth shows that the reported that maximum stability was attained at
bubble size distribution at a level 100 mm above the contact angles between 80j and 95j, while at contact
interface is closer to that at a level of 200 mm than the angles above about 95j, the foams were found to be
distribution just above the froth – liquid interface. destabilized drastically by the particles.
Bearing in mind that the size distributions given in The present study suggests that the wettability of
these figures are plotted at equal distance, this indi- the particles in the bubble film has a significant effect
cates that bubble coalescence occurs at higher rates in on the bubble size growth in the froth zone. Our
the lower part of the froth. results are in partial agreement with the results of
Dippenaar (1982a), Johansson and Pugh (1992) and
Aveyard et al. (1994). However, the conclusion that
4. Discussion particles with h>80j destroy the froth does not appear
to have been verified in our experiments. In the
The presence of solid particles is believed to have presence of particles with strong hydrophobicity,
a strong influence on froth stability. Dippenaar bubbles coalesce more rapidly than in the presence
(1982a) studied the mechanism of particle – film of the moderately hydrophobic glass, but less rapidly
interactions on a thin film. He found that spherical, than in the presence of the weakly hydrophobic
highly hydrophobic particles, especially spheres, with particles, as is evident in Figs. 2, 3 and 4. In the case
contact angles greater than 90j destabilize froth. of a low contact angle, where the force of attachment
The destabilization was the result of the thinning of of the particle to a film is low, perhaps drainage of
the inter-bubble liquid bridged by the particle. Par- liquid by gravity and capillary forces causes film
ticles with irregular shapes could rupture the films rupture. This indicates that the surface hydrophobicity
even when h < 90j. Johansson and Pugh (1992) of solid particles is not the only factor in determining
studied the influence of particles with varying hy- the coalescence process and that other factors should
drophobicity on froth stability. Both dynamic and also be taken into account.
static froth stability measurements, as well as micro- To understand the effect of hydrophobic particles
interferometric studies on thin aqueous films, were on flotation performance, the recovery of hydropho-
carried out with quartz particles having various size bic particles should be considered in conjunction with
factions and degrees of hydrophobicities. For the the bubble size distribution in the froth zone. Table 2
26– 44 Am size fraction, they found that hydropho- shows the recovery of glass particles (the hydropho-
bic particles could be transformed into very effective bic particles), water and the entrained solid (silica).
S. Ata et al. / Int. J. Miner. Process. 72 (2003) 255–266 263

Table 2 bicities, as used in this paper, were employed in the


Glass spheres (hydrophobic particles) recovery, water and entrained experimental work. The results showed that the
solid recovery rate obtained with three systems studied
strongly hydrophobic particles could be collected
System Contact Glass Water Silica
by rising bubbles at higher rates than the moderately
angle spheres recovery recovery
recovery rate rate and weakly hydrophobic particles, provided that the
(%) (g/min) (g/min) froth fluidity is high enough to allow particles to be
Sparger ‘1’a h = 82j 86 100 – transported to the launder. This indicates that attach-
(glass spheres) h = 66j 72 135 – ment process takes place as soon as particles impinge
h = 50j 42 68 – upon the surface of the bubbles. Thus, even though
Sparger ‘1’b h = 82j 84 164 12.1 the detachment rate of strongly hydrophobic glass in
(glass spheres + h = 66j 75 202 14.8
the froth may be higher due to bubble coalescence,
silica) h = 50j 51 100 6.3
Sparger ‘0’b h = 82j 70 66 4.5 they may easily attach again. When the froth contains
(glass spheres + h = 66j 61 80 5.1 particles of a high level of hydrophobicity, one may
silica) h = 50j 38 59 3.2 expect that the detachment and attachment processes
a
Froth depth: 320 mm. in the froth continue until the surface of the bubbles
b
Froth depth: 450 mm. is densely packed with the hydrophobic mineral.
Thus, it would be expected that this phenomenon is
Interestingly, strongly hydrophobic glass (contact more likely to occur only in a lightly mineralized
angle of 82j) gives the highest flotation recovery froth where there is sufficient bubble surface for
even though these particles are more capable of attachment.
rupturing bubble films than the moderately hydro- Table 3 shows the bubble size ratio in the froth (the
phobic glass in the froth. This observation applies to ratio of the Sauter-mean bubble diameter at a defined
all systems studied. The question now arises as to level and the Sauter-mean bubble diameter at the
why the highest recovery is observed with the strong- interface level) with Sparger ‘1’ and Sparger ‘0’ in
ly hydrophobic particles. An explanation may be the mixed system. The most obvious phenomenon
offered based on the high re-collection rate of par- here is that the bubbles experience more coalescence
ticles in the froth zone due to their high level of at the interface level.
hydrophobicity. In a simple gas– liquid froth, the stability of the
More recently, the authors conducted an experi- froth is directly related to the stability of the liquid
mental study on the collection rate of hydrophobic films separating two bubbles. As soon as the liquid
particles in the froth phase (Ata et al., 2002). Hy- drains out, a coalescence mechanism takes place. In a
drophobic particles were deliberately introduced into mineralized froth, mineral particles attaching to the
the froth phase. The flotation cell was designed so air – water interface or remaining in the film may
that froth and pulp phase could be readily separated change the properties of froth. In most cases, a
and collection of froth fed particles could be mea- mineralized froth exhibits a different behaviour from
sured directly. Glass particles of different hydropho- the two-phase froth. However, it is not clear which

Table 3
Bubble size ratio in the froth obtained with Sparger ‘1’ and Sparger ‘0’ in the mixed system
Distance from the Bubble diameter ratio (Sparger ‘0’) Bubble diameter ratio (Sparger ‘1’)
interface (mm)
h = 82j h = 66j h = 50j h = 82j h = 66j h = 50j
0 1.00 1.00 1.00 1.00 1.00 1.00
60 1.15 1.16 1.32 1.23 1.03 1.34
120 1.33 1.28 1.54 1.35 1.12 1.51
180 1.37 1.31 1.71 1.41 1.23 1.69
240 1.55 1.33 1.76 1.44 1.25 1.76
300 1.72 1.39 1.97 1.57 1.30 1.88
360 1.84 1.50 2.12 1.59 1.32 2.03
264 S. Ata et al. / Int. J. Miner. Process. 72 (2003) 255–266

process is the rate-determining step for bubble coa- lescence in the froth. However, if we assume that
lescence in a dynamic mineralized froth. particles should first enter the air – liquid interface and
Previous observations on the effect of particles on form bridges across the lamellae, in order to play a
the film rupture do not offer a unique conclusion. role in the lifetime of the bubble, then the upper level
Dippenaar (1982b) suggested that the rate-determin- of the froth where the bubble films are thinner would
ing step for film rupture is the thinning of the be the preferred area for this process. Results from
lamellae to a thickness near that of the particle size Figs. 2, 3 and 4 and Table 3, however, show that the
(required for bridging). Once particles bridge the rate of bubble coalescence is higher in the lower part
films, they destroy them. Frye and Berg (1989) of froth than in the upper level. This suggests that the
proposed a mechanism for particle-induced film physical processes governing the behaviour of two-
rupture and performed a hydrodynamic analysis to phase (i.e., gas and liquid) foams predominantly
determine criteria for effective antifoam action by govern the behaviour of mineralized froths (i.e., gas,
solid particles. They calculated the thinning time, liquid and solid) close to the pulp – froth interface. It is
which was compared with the rupture time, to worth mentioning that a similar observation was made
estimate the rate-determining step. For small particles in a two-phase column froth by Yianatos et al. (1986).
(b100 Am), rupture times were negligible and The interface is a region where bubbles begin to pack
thinning was the rate-determining step, in agreement together and arrange themselves to the new environ-
with the conclusion of Dippenaar (1982b). However, ment. Most of the liquid and solids is squeezed out
if the particle size was bigger than 100 Am, then due to this arrangement. This sudden change appears
thinning time declined and rupture became rate to cause bubble coalescence, and the bubble size rises
determining. sharply.
Regarding the effect of particle size and concen- The contribution of particles to the foam stability is
tration on the froth stability, Livshits and Dudenkov only possible when they are in the films. For this to
(1965) proposed that there is an optimum particle size occur, the film should be sufficiently thin. In the
range that promotes bubble coalescence. Coarse and present study, it is not possible to identify in which
very fine hydrophobic particles may not destroy froth part of the froth particles gives rise to film collapse or
because coarse particles act as buffers between two film survive. However, in general, if Figs. 2, 3 and 4
bubbles, thus slowing down bubble coalescence while are closely examined, it is seen that in the presence of
very fine particles would drain back with liquid to the weakly hydrophobic particles, the bubble destruction
pulp phase. Tao et al. (2000) showed that particles rate with respect the froth height is more rapid. This
< 150 Am destabilized froth at lower concentrations indicates that even the most hydrophobic particles
and stabilized it at higher concentrations while par- used in the present study add some stability to the
ticles < 30 Am always showed froth-breaking ability. froth phase, and the lifetime of the bubbles in the
The froth-destabilizing effect of fine hydrophobic presence of particles appears to be longer than those in
particles was attributed to the consumption of frother a particle-free system.
in the cell due to its adsorption on solids. Therefore, it Of special interest in this study is the investigation
is not clear whether the particles or the lack of frother of whether entrained solids affect the bubble growth
caused froth collapse in their system. The work by rate through the froth column. Evidently, the presence
Livshits and Dudenkov (1965) and Tao et al. (2000) of gangue mineral in the froth decreases the bubble
does not provide any direct information on which coalescence rate. A possible explanation is that they
process is responsible for bubble coalescence in a increase the viscosity of slurry retained between bub-
dynamic froth. However, it suggests that the charac- bles, which blocks the channels through the lamellae,
teristics of mineral particles (i.e., size, shape concen- preventing drainage. This finding suggests that bubble
tration) may be an important factor in deciding coalescence in the froth may be reduced by using an
whether film thinning or rupture is the rate-determin- appropriate frother type and concentration. It also
ing step. shows that the entrained solids may cause a persistent
It is apparent that there is no conclusive agreement froth, so selectivity may be a problem if the concen-
on which process primarily governs the bubble coa- tration of the fine gangue in the froth is high.
S. Ata et al. / Int. J. Miner. Process. 72 (2003) 255–266 265

An expression for the viscosity of a slurry has been The bubble growth rate is sensitive to the degree of
given by Chung and Adelman (1978): hydrophobicity of the particles present in the froth
zone. The rate of change in the Sauter-mean bubble
l* ¼ lð1 þ 2:5us þ 4:375u2s Þ ð2Þ diameter is highest with the weakly hydrophobic
particles and lowest in case of the intermediate hy-
drophobic particles. Taking bubble coalescence rate as
where l* is the viscosity of the slurry, l is the
an indication of froth stability, the results show that
viscosity of water and us is the volumetric fraction
maximum froth stability is attained when froth con-
of solids in the slurry and can be calculated by the
tains particles with moderate surface hydrophobicity.
following relationship:
The overall recovery, however, suggests that particles
with a strong level of floatability exhibit the highest
Xs
us ¼   ð3Þ flotation recovery probably as a result of particle
q reattachment in the froth. Thus, a high bubble coales-
Xs þ s
q cence rate in the froth zone is not necessarily associ-
ated with low flotation recovery.
where Xs is the solid concentration in the slurry and qs In this study, it was also found that the amount of
and q are the density of the solid and liquid, respec- entrained solids in the froth considerably reduces
tively. We can obtain an order of magnitude estimate bubble coalescence, probably by increasing the slurry
of the viscosity of the liquid in the froth by applying viscosity between the bubble films, and reducing the
these equations to the relevant slurries, which may drainage rate of liquid films, and possibly by mechan-
become entrained into the froth. ical blocking.
Using Eqs. (2) and (3), the slurry viscosity is found
to be 0.00108 and 0.00116 Pa s for glass spheres alone
and mixed systems (glass spheres + silica), respective- Acknowledgements
ly. The increase in the effective viscosity will reduce
the drainage rate of the liquid, thereby reducing the Seher Ata would like to acknowledge financial
coalescence of bubbles in the froth. support from the University of Newcastle. The authors
acknowledge the support of the Australian Research
Council in funding the Centre for Multiphase Pro-
5. Conclusions cesses, under its Special Research Centre Program.

The bubble size as a function of height above the


froth – pulp interface has been studied closely through References
photography in a specially designed continuous labo-
ratory cell. Special attention was paid to the effect of Ata, S., Ahmed, N., Jameson, G.J., 2002. Collection of hydro-
phobic particles in the froth phase. Int. J. Miner. Process. 64,
surface hydrophobicity of particles, entrained solid 101 – 122.
and initial bubble size (bubble size in the pulp zone) Aveyard, R., Binks, B.P., Fletcher, P.D.I., Peck, T.G., Rutherford,
on the sizes of bubbles in the froth. In order to assess C.E., 1994. Aspects of aqueous foam stability in the presence of
the real effect of hydrophobicity on the froth stability, hydrocarbon oils and solid particles. Adv. Colloid Interface Sci.
48, 93 – 120.
particles with a range of contact angles (50j, 66j and
Chung, Y.M., Adelman, S.A., 1978. Transport properties of con-
82j) and hydrophilic solid were used in the experi- centrated polymer solutions. Hydrodynamic mean field theory
mental programme. of the viscosity of a sphere suspension. J. Chem. Phys. 69,
The results show that the Sauter-mean bubble 3146 – 3149.
diameter in the froth increases with increasing height Dippenaar, A., 1982a. The destabilization of froth by solids: I. The
above the froth –pulp interface, which indicates that mechanism of film rupture. Int. J. Miner. Process. 9, 1 – 14.
Dippenaar, A., 1982b. The destabilisation of froth by solids: II. The
bubble coalescence occurs at all levels of the froth. rate-determining step. Int. J. Miner. Process. 9, 15 – 27.
The change in bubble size is essentially strong close Frye, G.C., Berg, J.C., 1989. Antifoam action by solid particles.
to the interface. J. Colloid Interface Sci. 127, 222 – 238.
266 S. Ata et al. / Int. J. Miner. Process. 72 (2003) 255–266

Johansson, G., Pugh, R.J., 1992. The influence of particle size and Tao, D., Luttrell, G.H., Yoon, R.-H., 2000. A parametric study of
hydrophobicity on the stability of mineralised froths. Int. J. froth stability and its effect on column flotation of fine particles.
Miner. Process. 34, 1 – 20. Int. J. Miner. Process. 59, 25 – 43.
Livshits, A.K., Dudenkov, S.V., 1965. Some factors in flotation Yianatos, J.B., Finch, J.A., Laplante, A.R., 1986. Holdup profile
froth stability. In: Arbiter, N. (Ed.), Proc. VII Int. Min. Proc. and bubble size distribution of flotation column froths. Can.
Cong. Gordon and Breach, New York, pp. 367 – 371. Metall. Q. 25, 23 – 29.

You might also like