Sie sind auf Seite 1von 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
Despite the importance of froth in the flotation
process, little attention has been paid to the details of
the mechanisms, which governs the bubble behaviour
in the froth column. The process is very difficult to
study because the presence of particles in the froth
column prevents the measurement of system parameters, such as gas holdup or even bubble size, accurately. As a result, attempts to model the flotation froth at a
macroscopic level rely much more on the knowledge
of two-phase froths rather than actual phenomena.

* Corresponding author. Tel.: +61-2-4921-6181; fax: +61-24960-1445.


E-mail address: Seher.Ata@newcastle.edu.au (S. Ata).
0301-7516/$ - see front matter D 2003 Elsevier B.V. All rights reserved.
doi:10.1016/S0301-7516(03)00103-0

However, a flotation froth can exhibit a quite different


structure from a two-phase foam since it contains
particles that in turn may play a key role in bubble
coalescence and froth stability. The behaviour of the
flotation froth therefore is governed not only by the
factors that rule two-phase foams such as film thinning
or gravity drainage but also by contributions from the
particles. The study of each process is equally important in understanding the mineralized froth.
The aim of the work reported herein is to obtain a
better understanding of froth behaviour under different conditions, through observation of the changes of
the bubble size distribution in the rising froth due to
coalescence. Special attention has been given to
particle hydrophobicity. The parameters of interest
are the hydrophobicity of the solid particles, the initial
bubble size and the concentration of gangue mineral.

256

S. Ata et al. / Int. J. Miner. Process. 72 (2003) 255266

2. Experimental
2.1. Materials
2.1.1. Silica
Washed quartz obtained from Commercial Minerals, Australia, under the brand name Silica #400 was
used as the hydrophilic (gangue) solid. The average
particle size of the sample (d50), determined using a
Malvern Particle Sizer, was 5 Am.
2.1.2. Glass spheres
Glass spheres were used as the hydrophobic model
system. The glass particles used in this study were
Mill Spec-13 grade supplied by Godfrey Blast Clean-

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


Am; the d80 was 82 Am.
Glass particles of different hydrophobicities (as
reflected in the contact angle) were prepared using
controlled silanation. Contact angles (h) chosen were
50j, 66j and 82j. Details of the preparation of glass
particles may be found in Ata et al. (2002).
2.2. The flotation apparatus
A schematic of the cell is shown in Fig. 1. The
flotation cell consists of a 2.3-l, 150-mm diameter
stirred vessel connected to a 50-mm diameter column
through a tapered transition piece. The transition piece
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) 255266

froth over wide operating conditions. The collection


zone was constructed from Perspex, and the column
section was made of glass to which bubbles do not
adhere. Bubbles were generated by introducing nitrogen through a sintered glass frit incorporated into the
base of the cell. The bubble size was varied by using
frits with different pore sizes.
The flotation cell was operated in a continuous
mode. A batch of feed was stirred in a 25-l vessel for
8 min and then fed into the cell through a constant head
tank. A constant flow rate of slurry through the cell
was achieved by the use of a peristaltic pump, which
was set to deliver the required flow of 650 ml/min. The
tailing stream was passed through a gravity overflow.
The gravity overflow was also used to maintain the
position of the froth liquid interface at a preset level in
the column. The superficial gas velocity was kept
constant at 1.2 cm/s, and frother Dowfroth 250, at a
constant dosage of 30 mg/l, was the only reagent added
to the cell during the experiments.
In the present study, the bubble size was determined by a photographic technique. Bubbles near the
column wall in the froth zone were photographed
using a Nikon camera after the concentrate reached a
steady state. Distortion of images due to the cylindrical surface of the column was eliminated by
employing a transparent viewing box filled with
water around the column. The box could be moved
up and down easily to allow bubbles to be photographed at various levels in the froth. A graduated
scale was secured to the front of the box to provide
the necessary scale. Tests indicated that side lighting
by the flash gave the best definition. The flash was
covered by a triple layer of tracing paper. The
camera setting was f/8 and 1/100 s shutter speed.
Black-and-white Agfa 25 ISO films were used to
provide extra definition. At least 500 bubbles were
sized in each case. The negatives of the photographs
were projected using an enlarger directly onto a
digitiser pad (Summagraphics) and sized using a
personal computer.

3. Results
In this study, the Sauter-mean diameter is used as
the representative size for the bubble size distribution
in the froth phase. The Sauter-mean bubble diameter

257

(d32) is defined as the volume-to-surface mean bubble


diameter:
P 3
ni d
d32 P i2
1
ni di
where ni is the number of bubbles with diameter di.
The Sauter-mean bubble diameter is commonly used
in flotation studies. This usage arises because the rate
of flotation is closely linked to the surface area of the
bubbles.
The axial profiles of the Sauter-mean diameter
(d32) through the bubble bed in the presence of solids
of different hydrophobicities are shown in Fig. 2.
Three levels of hydrophobicity were investigated,
i.e., strongly, intermediate and weakly corresponding
to contact angles of 82j, 66j and 50j, respectively.
Tests were conducted at 1.2 cm/s superficial gas
velocity; 320 mm of froth depth; 30 ppm of frother
concentration; 650 cm3/min feed flow rate; and 7.5%
solids concentration. It should be noted that, except
for the parameter under study, these experimental
conditions were kept unchanged in all of the tests.
However, under the various operating conditions, it
was observed that the froths formed above the pulp
zone were generally different in height. Thus, for each
system studied, the froth depth was adjusted to allow a
free flow of froth over the cell lip.
It is evident from Fig. 2 that the Sauter-mean bubble
diameter, in the presence of hydrophobic particles,
increases with increasing distance from the pulp froth
interface. The major increase in d32 occurs with
particles of the lowest contact angle; these particles
appear to have less effect on the stability of the froth.
Interestingly, the particles that have greatest bubble
stability were those of intermediate hydrophobicity.
In order to observe the effect of entrained particles
on the bubble size in the froth phase, a series of tests
were conducted with feeds containing a mixture of
hydrophobic glass spheres and gangue mineral (silica). The results are shown in Fig. 3 where the Sautermean bubble diameter is plotted as a function of the
height above the pulp froth interface for particles of
the three hydrophobicities. The total concentration of
solids in the feed was 15% (w/w), with equal masses
of glass and silica; the froth depth was at 450 mm.
The remaining operating conditions are as given for
Fig. 2. Silica did not exhibit any floatability, so it was

258

S. Ata et al. / Int. J. Miner. Process. 72 (2003) 255266

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


measurement was also carried out in the system where
only hydrophilic particles (silica) were present to
determine whether the hydrophobic glass particles
do in fact play a role in all cases. The concentration

of silica in the flotation cell for these tests was 7.5%


(w/w). Results are given in Fig. 3.
It is seen that the d32 values obtained with the
mixed system show a trend similar to that observed
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) 255266


Table 1
Statistical analysis of bubble size in the mixed and glass spheres
systems for Sparger 1 bubbles
Contact angle (h)

P value

h = 82j
h = 66j
h = 50j

0.0026
0.0203
0.2580

are present. The d32 increases continuously with froth


height for all three hydrophobicities and hydrophilic
particles. Again, it is significant to observe that the
increase in d32 is the lowest for intermediate hydrophobic particles, while being the highest for weakly
hydrophobic particles and hydrophilic silica. Note that
the bubble growth rate in the presence of hydrophilic
particles and weakly hydrophobic glass seems to be
similar, suggesting that the contribution of weakly
hydrophobic glass to the stability of the froth is
minimal.
A comparison between Figs. 2 and 3 shows that the
change in the Sauter-mean bubble diameter with
respect to the froth height is considerably reduced in
the mixed system for all three hydrophobicities. For
example, in a system of hydrophobic glass particles
alone, at a froth level of 200 mm, the d32 for the
strongly, intermediate and weakly hydrophobic particles is 1.6, 1.5 and 1.9 mm, respectively, while the

259

corresponding values in the mixed system are 1.1, 0.9


and 1.3 mm, respectively.
Table 1 gives a statistical analysis of bubble size at
the interface in the two-solid (glass spheres and silica)
and the one-solid (glass spheres) systems for the three
hydrophobic particles, using the F-test. Probability
( P) values more than or equal to 0.05 were indicative
of the size of the bubbles in two systems being
statistically different at a confidence level z 95%. It
is seen that with strongly and intermediate hydrophobic glass, the bubble size at the interface with glass
spheres and mixed systems is the same at the 95%
confidence level, while with weakly hydrophobic
particles, the size of the bubbles at the pulp froth
interface is different in the two systems.
Tests were also carried out to investigate whether
the bubble size in the froth zone depends on the initial
bubble size (the bubble size in the liquid phase). In the
present study, the initial bubble size was varied
independently using another sparger having a different porosity. For convenience, this will be designated
as Sparger 0. Similarly, the sparger used in Figs. 2
and 3 will be nominated as Sparger 1. Sparger 0
has a larger pore size distribution than Sparger 1,
which means that under the same experimental conditions, it produces larger bubbles. The comparison of
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) 255266

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) 255266

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


present, to eliminate the effect of particle hydrophobicity 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 substantial 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 Sautermean 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, respectively. 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.

261

262

S. Ata et al. / Int. J. Miner. Process. 72 (2003) 255266

seen from the figures that the Sauter-mean bubble


diameter increases for each with increasing distance
from the pulp froth interface for all the hydrophobic
particles. The distribution at the interface has a very
narrow size range. As the froth height above the
interface increases, the bubble size distribution
becomes wider, suggesting that large bubbles emerge,
but also that a significant number of small bubbles
remain. This may be seen clearly in Fig. 6a c where
bubbles were photographed at different levels in the
froth. The interesting aspect of these pictures lies in
the fact that the structure of the froth is very different
from the froth structure that is generally used in
flotation froth modeling. Even at the highest level of
the froth (Fig. 6a) where the liquid content is the
lowest, the shape of the plateau border is not defined
clearly.
In Fig. 5a c, a comparison between the bubble
size distributions at three levels of froth shows that the
bubble size distribution at a level 100 mm above the
interface is closer to that at a level of 200 mm than the
distribution just above the froth liquid interface.
Bearing in mind that the size distributions given in
these figures are plotted at equal distance, this indicates that bubble coalescence occurs at higher rates in
the lower part of the froth.

4. Discussion
The presence of solid particles is believed to have
a strong influence on froth stability. Dippenaar
(1982a) studied the mechanism of particle film
interactions on a thin film. He found that spherical,
highly hydrophobic particles, especially spheres, with
contact angles greater than 90j destabilize froth.
The destabilization was the result of the thinning of
the inter-bubble liquid bridged by the particle. Particles with irregular shapes could rupture the films
even when h < 90j. Johansson and Pugh (1992)
studied the influence of particles with varying hydrophobicity on froth stability. Both dynamic and
static froth stability measurements, as well as microinterferometric studies on thin aqueous films, were
carried out with quartz particles having various size
factions and degrees of hydrophobicities. For the
26 44 Am size fraction, they found that hydrophobic particles could be transformed into very effective

film breakers if they have a contact angle of more


than 80j. With particles of intermediate hydrophobicity (corresponding to h = 65j), the froth stability
was maximized suggesting that there is an optimum
particle hydrophobicity that promotes froth stability.
From observations on thin films containing particles,
it was suggested that particles of intermediate hydrophobicity were capable of forming stable bridges
across the film, increasing the rigidity of the froth
structure. Particles of high hydrophobicity (h = 82j)
penetrated the interface to a much greater extent and
ruptured the film, thus leading to unstable film
bridging. Particles with low degree of hydrophobicity
corresponding to a contact angle of h < 40j were
found to stream out into the lamella and did not
contribute to the stability of the froth film. Aveyard et
al. (1994) studied the effect of spherical glass beads
of varying hydrophobicity on foam stability, and
reported that maximum stability was attained at
contact angles between 80j and 95j, while at contact
angles above about 95j, the foams were found to be
destabilized drastically by the particles.
The present study suggests that the wettability of
the particles in the bubble film has a significant effect
on the bubble size growth in the froth zone. Our
results are in partial agreement with the results of
Dippenaar (1982a), Johansson and Pugh (1992) and
Aveyard et al. (1994). However, the conclusion that
particles with h>80j destroy the froth does not appear
to have been verified in our experiments. In the
presence of particles with strong hydrophobicity,
bubbles coalesce more rapidly than in the presence
of the moderately hydrophobic glass, but less rapidly
than in the presence of the weakly hydrophobic
particles, as is evident in Figs. 2, 3 and 4. In the case
of a low contact angle, where the force of attachment
of the particle to a film is low, perhaps drainage of
liquid by gravity and capillary forces causes film
rupture. This indicates that the surface hydrophobicity
of solid particles is not the only factor in determining
the coalescence process and that other factors should
also be taken into account.
To understand the effect of hydrophobic particles
on flotation performance, the recovery of hydrophobic particles should be considered in conjunction with
the bubble size distribution in the froth zone. Table 2
shows the recovery of glass particles (the hydrophobic particles), water and the entrained solid (silica).

S. Ata et al. / Int. J. Miner. Process. 72 (2003) 255266


Table 2
Glass spheres (hydrophobic particles) recovery, water and entrained
solid recovery rate obtained with three systems studied
System

Contact
angle

Glass
spheres
recovery
(%)

Water
recovery
rate
(g/min)

Silica
recovery
rate
(g/min)

Sparger 1a
(glass spheres)

h = 82j
h = 66j
h = 50j
h = 82j
h = 66j
h = 50j
h = 82j
h = 66j
h = 50j

86
72
42
84
75
51
70
61
38

100
135
68
164
202
100
66
80
59

12.1
14.8
6.3
4.5
5.1
3.2

Sparger 1b
(glass spheres +
silica)
Sparger 0b
(glass spheres +
silica)
a
b

Froth depth: 320 mm.


Froth depth: 450 mm.

Interestingly, strongly hydrophobic glass (contact


angle of 82j) gives the highest flotation recovery
even though these particles are more capable of
rupturing bubble films than the moderately hydrophobic glass in the froth. This observation applies to
all systems studied. The question now arises as to
why the highest recovery is observed with the strongly hydrophobic particles. An explanation may be
offered based on the high re-collection rate of particles in the froth zone due to their high level of
hydrophobicity.
More recently, the authors conducted an experimental study on the collection rate of hydrophobic
particles in the froth phase (Ata et al., 2002). Hydrophobic particles were deliberately introduced into
the froth phase. The flotation cell was designed so
that froth and pulp phase could be readily separated
and collection of froth fed particles could be measured directly. Glass particles of different hydropho-

263

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


experimental work. The results showed that the
strongly hydrophobic particles could be collected
by rising bubbles at higher rates than the moderately
and weakly hydrophobic particles, provided that the
froth fluidity is high enough to allow particles to be
transported to the launder. This indicates that attachment process takes place as soon as particles impinge
upon the surface of the bubbles. Thus, even though
the detachment rate of strongly hydrophobic glass in
the froth may be higher due to bubble coalescence,
they may easily attach again. When the froth contains
particles of a high level of hydrophobicity, one may
expect that the detachment and attachment processes
in the froth continue until the surface of the bubbles
is densely packed with the hydrophobic mineral.
Thus, it would be expected that this phenomenon is
more likely to occur only in a lightly mineralized
froth where there is sufficient bubble surface for
attachment.
Table 3 shows the bubble size ratio in the froth (the
ratio of the Sauter-mean bubble diameter at a defined
level and the Sauter-mean bubble diameter at the
interface level) with Sparger 1 and Sparger 0 in
the mixed system. The most obvious phenomenon
here is that the bubbles experience more coalescence
at the interface level.
In a simple gas liquid froth, the stability of the
froth is directly related to the stability of the liquid
films separating two bubbles. As soon as the liquid
drains out, a coalescence mechanism takes place. In a
mineralized froth, mineral particles attaching to the
air water interface or remaining in the film may
change the properties of froth. In most cases, a
mineralized froth exhibits a different behaviour from
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
interface (mm)

Bubble diameter ratio (Sparger 0)

Bubble diameter ratio (Sparger 1)

h = 82j

h = 66j

h = 50j

h = 82j

h = 66j

h = 50j

0
60
120
180
240
300
360

1.00
1.15
1.33
1.37
1.55
1.72
1.84

1.00
1.16
1.28
1.31
1.33
1.39
1.50

1.00
1.32
1.54
1.71
1.76
1.97
2.12

1.00
1.23
1.35
1.41
1.44
1.57
1.59

1.00
1.03
1.12
1.23
1.25
1.30
1.32

1.00
1.34
1.51
1.69
1.76
1.88
2.03

264

S. Ata et al. / Int. J. Miner. Process. 72 (2003) 255266

process is the rate-determining step for bubble coalescence in a dynamic mineralized froth.
Previous observations on the effect of particles on
the film rupture do not offer a unique conclusion.
Dippenaar (1982b) suggested that the rate-determining step for film rupture is the thinning of the
lamellae to a thickness near that of the particle size
(required for bridging). Once particles bridge the
films, they destroy them. Frye and Berg (1989)
proposed a mechanism for particle-induced film
rupture and performed a hydrodynamic analysis to
determine criteria for effective antifoam action by
solid particles. They calculated the thinning time,
which was compared with the rupture time, to
estimate the rate-determining step. For small particles
(b100 Am), rupture times were negligible and
thinning was the rate-determining step, in agreement
with the conclusion of Dippenaar (1982b). However,
if the particle size was bigger than 100 Am, then
thinning time declined and rupture became rate
determining.
Regarding the effect of particle size and concentration on the froth stability, Livshits and Dudenkov
(1965) proposed that there is an optimum particle size
range that promotes bubble coalescence. Coarse and
very fine hydrophobic particles may not destroy froth
because coarse particles act as buffers between two
bubbles, thus slowing down bubble coalescence while
very fine particles would drain back with liquid to the
pulp phase. Tao et al. (2000) showed that particles
< 150 Am destabilized froth at lower concentrations
and stabilized it at higher concentrations while particles < 30 Am always showed froth-breaking ability.
The froth-destabilizing effect of fine hydrophobic
particles was attributed to the consumption of frother
in the cell due to its adsorption on solids. Therefore, it
is not clear whether the particles or the lack of frother
caused froth collapse in their system. The work by
Livshits and Dudenkov (1965) and Tao et al. (2000)
does not provide any direct information on which
process is responsible for bubble coalescence in a
dynamic froth. However, it suggests that the characteristics of mineral particles (i.e., size, shape concentration) may be an important factor in deciding
whether film thinning or rupture is the rate-determining step.
It is apparent that there is no conclusive agreement
on which process primarily governs the bubble coa-

lescence in the froth. However, if we assume that


particles should first enter the air liquid interface and
form bridges across the lamellae, in order to play a
role in the lifetime of the bubble, then the upper level
of the froth where the bubble films are thinner would
be the preferred area for this process. Results from
Figs. 2, 3 and 4 and Table 3, however, show that the
rate of bubble coalescence is higher in the lower part
of froth than in the upper level. This suggests that the
physical processes governing the behaviour of twophase (i.e., gas and liquid) foams predominantly
govern the behaviour of mineralized froths (i.e., gas,
liquid and solid) close to the pulp froth interface. It is
worth mentioning that a similar observation was made
in a two-phase column froth by Yianatos et al. (1986).
The interface is a region where bubbles begin to pack
together and arrange themselves to the new environment. Most of the liquid and solids is squeezed out
due to this arrangement. This sudden change appears
to cause bubble coalescence, and the bubble size rises
sharply.
The contribution of particles to the foam stability is
only possible when they are in the films. For this to
occur, the film should be sufficiently thin. In the
present study, it is not possible to identify in which
part of the froth particles gives rise to film collapse or
film survive. However, in general, if Figs. 2, 3 and 4
are closely examined, it is seen that in the presence of
weakly hydrophobic particles, the bubble destruction
rate with respect the froth height is more rapid. This
indicates that even the most hydrophobic particles
used in the present study add some stability to the
froth phase, and the lifetime of the bubbles in the
presence of particles appears to be longer than those in
a particle-free system.
Of special interest in this study is the investigation
of whether entrained solids affect the bubble growth
rate through the froth column. Evidently, the presence
of gangue mineral in the froth decreases the bubble
coalescence rate. A possible explanation is that they
increase the viscosity of slurry retained between bubbles, which blocks the channels through the lamellae,
preventing drainage. This finding suggests that bubble
coalescence in the froth may be reduced by using an
appropriate frother type and concentration. It also
shows that the entrained solids may cause a persistent
froth, so selectivity may be a problem if the concentration of the fine gangue in the froth is high.

S. Ata et al. / Int. J. Miner. Process. 72 (2003) 255266

An expression for the viscosity of a slurry has been


given by Chung and Adelman (1978):
l* l1 2:5us 4:375u2s

where l* is the viscosity of the slurry, l is the


viscosity of water and us is the volumetric fraction
of solids in the slurry and can be calculated by the
following relationship:
Xs

us 
q
Xs s
q

where Xs is the solid concentration in the slurry and qs


and q are the density of the solid and liquid, respectively. We can obtain an order of magnitude estimate
of the viscosity of the liquid in the froth by applying
these equations to the relevant slurries, which may
become entrained into the froth.
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), respectively. The increase in the effective viscosity will reduce
the drainage rate of the liquid, thereby reducing the
coalescence of bubbles in the froth.

5. Conclusions
The bubble size as a function of height above the
froth pulp interface has been studied closely through
photography in a specially designed continuous laboratory cell. Special attention was paid to the effect of
surface hydrophobicity of particles, entrained solid
and initial bubble size (bubble size in the pulp zone)
on the sizes of bubbles in the froth. In order to assess
the real effect of hydrophobicity on the froth stability,
particles with a range of contact angles (50j, 66j and
82j) and hydrophilic solid were used in the experimental programme.
The results show that the Sauter-mean bubble
diameter in the froth increases with increasing height
above the froth pulp interface, which indicates that
bubble coalescence occurs at all levels of the froth.
The change in bubble size is essentially strong close
to the interface.

265

The bubble growth rate is sensitive to the degree of


hydrophobicity of the particles present in the froth
zone. The rate of change in the Sauter-mean bubble
diameter is highest with the weakly hydrophobic
particles and lowest in case of the intermediate hydrophobic particles. Taking bubble coalescence rate as
an indication of froth stability, the results show that
maximum froth stability is attained when froth contains particles with moderate surface hydrophobicity.
The overall recovery, however, suggests that particles
with a strong level of floatability exhibit the highest
flotation recovery probably as a result of particle
reattachment in the froth. Thus, a high bubble coalescence rate in the froth zone is not necessarily associated with low flotation recovery.
In this study, it was also found that the amount of
entrained solids in the froth considerably reduces
bubble coalescence, probably by increasing the slurry
viscosity between the bubble films, and reducing the
drainage rate of liquid films, and possibly by mechanical blocking.

Acknowledgements
Seher Ata would like to acknowledge financial
support from the University of Newcastle. The authors
acknowledge the support of the Australian Research
Council in funding the Centre for Multiphase Processes, under its Special Research Centre Program.

References
Ata, S., Ahmed, N., Jameson, G.J., 2002. Collection of hydrophobic particles in the froth phase. Int. J. Miner. Process. 64,
101 122.
Aveyard, R., Binks, B.P., Fletcher, P.D.I., Peck, T.G., Rutherford,
C.E., 1994. Aspects of aqueous foam stability in the presence of
hydrocarbon oils and solid particles. Adv. Colloid Interface Sci.
48, 93 120.
Chung, Y.M., Adelman, S.A., 1978. Transport properties of concentrated polymer solutions. Hydrodynamic mean field theory
of the viscosity of a sphere suspension. J. Chem. Phys. 69,
3146 3149.
Dippenaar, A., 1982a. The destabilization of froth by solids: I. The
mechanism of film rupture. Int. J. Miner. Process. 9, 1 14.
Dippenaar, A., 1982b. The destabilisation of froth by solids: II. The
rate-determining step. Int. J. Miner. Process. 9, 15 27.
Frye, G.C., Berg, J.C., 1989. Antifoam action by solid particles.
J. Colloid Interface Sci. 127, 222 238.

266

S. Ata et al. / Int. J. Miner. Process. 72 (2003) 255266

Johansson, G., Pugh, R.J., 1992. The influence of particle size and
hydrophobicity on the stability of mineralised froths. Int. J.
Miner. Process. 34, 1 20.
Livshits, A.K., Dudenkov, S.V., 1965. Some factors in flotation
froth stability. In: Arbiter, N. (Ed.), Proc. VII Int. Min. Proc.
Cong. Gordon and Breach, New York, pp. 367 371.

Tao, D., Luttrell, G.H., Yoon, R.-H., 2000. A parametric study of


froth stability and its effect on column flotation of fine particles.
Int. J. Miner. Process. 59, 25 43.
Yianatos, J.B., Finch, J.A., Laplante, A.R., 1986. Holdup profile
and bubble size distribution of flotation column froths. Can.
Metall. Q. 25, 23 29.

Das könnte Ihnen auch gefallen