Download as pdf or txt
Download as pdf or txt
You are on page 1of 32

Int. J. Miner. Process. 56 Ž1999.

133–164

Bubble–particle attachment and detachment in


flotation
John Ralston ) , Daniel Fornasiero, Robert Hayes
Ian Wark Research Institute, UniÕersity of South Australia, The LeÕels Campus, Mawson Lakes, Adelaide,
SA 5095, Australia

Received 3 June 1998; received in revised form 12 October 1998; accepted 27 November 1998

Abstract

The mechanism by which particles and bubbles interact captures many of the central concepts
of colloid science and hydrodynamics and is an example of heterocoagulation. Hydrodynamics,
interfacial Žincluding capillary. forces, particle and bubble behaviour and solution chemistry are
all interwoven. The processes of attachment and detachment are focused upon here. We deal with
the identification of a flotation ‘domain’, the deformation of a bubble surface upon interaction
with a solid surface, the kinetics of three phase contact line expansion and the determination of
attachment efficiencies through to the direct measurement of bubble–particle interaction forces.
The results, concepts and implications of this work are discussed. q 1999 Elsevier Science B.V.
All rights reserved.

Keywords: bubble–particle; attachment; detachment

1. Introduction

The capture of particles by rising bubbles is the central process in froth flotation. For
efficient capture to occur between a bubble and a hydrophobic particle, they must first
undergo a sufficiently close encounter, a process which is controlled by the hydrody-
namics governing their approach in the aqueous environment in which they are normally
immersed. Should they approach quite closely within the range of attractive surface
forces, the intervening liquid film between the bubble and the particle will drain, leading

)
Corresponding author. Fax: q61-8-302-3683; E-mail: john.ralston@unisa.edu.au

0301-7516r99r$ - see front matter q 1999 Elsevier Science B.V. All rights reserved.
PII: S 0 3 0 1 - 7 5 1 6 Ž 9 8 . 0 0 0 4 6 - 5
134
Table 1
Key papers in understanding fundamental flotation substeps
1948 A fundamental paper on the kinetics of the flotation process appeared Žby Sutherland, 1948. in Australia. This invoked induction time,
described particle size effects in flotation and catalysed other similar approaches. While it was preceded by other efforts, it was the first
comprehensive effort to describe recovery-size-time data in a fundamental manner.

J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164


1960–1961 In Moscow, Derjaguin and Dukhin Ž1960. produced a key paper on the theory of flotation of small and medium-size particles.
Hydrodynamics, surface forces and diffusiophoresis were all used in this theory. This seminal work resulted in an acceleration of
fundamental flotation research worldwide.
1972 Blake and Kitchener Ž1972., working together in London, published some very careful measurements of the thickness of aqueous
films on hydrophobic quartz surfaces. Film thicknesses, measured as a function of salt concentration, were shown to depend on the
electrical double layer force. Film instability occurred on hydrophobic surfaces at film thicknesses less than about 60 nm. This value,
which was smaller than the range of the electrical double layer force, represented the combined effects of hydrophobic force, surface
heterogeneities and external disturbances. It hinted at the length dependence of hydrophobic forces, information which was subsequently
determined by surface force experiments after 1982.
1976 Scheludko Ž1967. and Scheludko et al. Ž1976. in Bulgaria considered how particles might become attached to a liquid surface and
developed the capillary theory of flotation.
1977 Anfruns and Kitchener Ž1986. published the first measurements of the absolute rate of capture of small particles in flotation. This was
the first critical test of collision theory under conditions where the bubble and particle surface chemistry was characterised and controlled.
1983 Schulze Ž1984. published a key textbook on the physico-chemical substeps which are important in flotation, drawing on a wide range
of hydrodynamic, surface chemical and engineering information.
circa 1980–present There has been a strong interest in developing reliable collision models ŽDobby and Finch, 1986; Dobby and Finch, 1987; Yoon and
Luttrell, 1989; Dai et al., 1998.. The surface force apparatus and, recently, the atomic force microscope colloid probe technique, have
provided very useful insight into electrical double layer, van der Waals and hydrophobic forces ŽChristenson and Claesson, 1988; Butt,
1994; Ducker et al., 1994; Wood and Sharma, 1995; Fielden et al., 1996.. Thin film drainage has been investigated between a rigid and
a deformable interface ŽLin and Slattery, 1982; Chen, 1984; Hewitt et al., 1993.. Attachment efficiencies have been measured. Reliable
methods for measuring contact angles on particles have been developed ŽCrawford et al., 1987; Diggins et al., 1990.. Major theoretical
and experimental advances in describing dynamic contact angles on well defined surfaces have been made Že.g., Blake, 1993; Hayes and
Ralston, 1993..
J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164 135

to a critical thickness at which rupture occurs. This is then followed by movement of the
three phase contact line Žthe boundary between the solid particle surface, receding liquid
phase and advancing gas phase. until a stable wetting perimeter is established. This
sequence of drainage, rupture and contact line movement constitutes the second process
of attachment. A stable particle–bubble union is thus formed. The particle may only be
dislodged from this state if it is supplied with sufficient kinetic energy to equal or
exceed the detachment energy, i.e., a third process of detachment can occur.
The capture Žor collection. efficiency E of a bubble and a particle may be defined as

E s Ec P E a P E s Ž 1.
where Ec is the collision efficiency, Ea is the attachment efficiency and Es is the
stability efficiency of the bubble–particle aggregate. This dissection of capture effi-
ciency into three process efficiencies was used by Derjaguin and Dukhin Ž1960. and
focuses attention on the three zones of bubble–particle capture where, in order,
hydrodynamic interactions, interfacial forces and bubble–particle aggregate stability are
dominant, as illustrated in Fig. 1. We should note that these zones are not completely
discrete, rather they grade into one another. Since surface forces are relatively short
range, they do not have any significant influence on the collision step. This is why
collision and attachment may be treated as essentially independent processes.
Our knowledge of the various efficiencies has been enhanced by a continuous, strong
research thrust, catalysed by Sutherland Ž1948. in 1948. Six of the most important
publications, described in Table 1, covering the period from 1948 to 1983, signalled
major advances in the field and have directly led to the present high level of research
activity. Two of the most important substeps in the froth flotation process, attachment
and detachment, are described below.

Fig. 1. Hydrodynamic Ž1., diffusiophoretic Ž2. and surface force Ž3. zones of interaction between a bubble and
a particle. ŽFrom Derjaguin and Dukhin, 1960, with permission..
136 J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164

2. Attachment

2.1. Zones of influence

Derjaguin and Dukhin Ž1960. identified zone 2 in Fig. 1 as that region where
diffusion effects are important. A strong electric field exists in this zone, for the liquid
flow around the moving bubble gives rise to a tangential stream at its surface which
destroys the equilibrium distribution of adsorbed ions there. Where surfactant is present,
it is continually swept from the upper to the lower surface of the bubble. Transport of
ionic surfactant to the moving bubble surface therefore takes place, leading to the
establishment of a concentration gradient. A strong electric field of order 3000 V cmy1
is established when the cation and anion diffusion coefficients differ, as they generally
do. Hence, charged particles entering zone 2 will experience an electrophoretic force in
precisely the same way as in an electrophoresis cell and would be either attracted
towards, or repelled from the bubble surface. Derjaguin and Dukhin coined the term
‘diffusiophoresis’ for this phenomenon, i.e., the ‘diffusiophoretic force’ therefore acts on
the particle as an additional force. To date, however, evidence confirming the presence
or absence of diffusiophoresis in flotation is equivocal and sparse. Apart from noting its
possible contribution to capture efficiency it is not pursued further here.
In zone 3, surface forces predominate once the thin film between the bubble and the
particle is reduced much below a few hundred nanometer. These forces can accelerate,
retard or even prevent the thinning of the liquid film between the particle and the
bubble. From a thermodynamic point of view, the free energy of a liquid film differs
from the bulk phase from which it is formed. This excess free energy was originally
called the ‘wedging apart’ or ‘disjoining’ pressure by Derjaguin and represents the
difference between the pressure within the film, p f , and that in the bulk liquid adjacent
to the solid surface, p l. Note that for a bubble pushed against a flat solid surface,
immersed in water, p b , the pressure within the bubble, is equal to p f . The Derjaguin and
Scheludko schools in Moscow and Sofia ŽScheludko, 1967; Derjaguin and Churaev,
1976. performed experimental measurements of disjoining pressures, providing both the
first real tests of the DLVO theory of surface forces, as well as the first accurate
experimental estimates of the Hamaker constant, respectively. The disjoining pressure
Žp . depends on the film thickness, h, and:
p Ž h. s pf y pl . Ž 2.
For mechanical equilibrium in a stable film, p Ž h. ) 0 and dprd h - 0.
If the liquid film is stable at all thicknesses, the liquid is said to completely wet the
solid. This occurs, for example, when an air bubble approaches a clean silica surface
immersed in water—in this instance, the Hamaker constant is negative and the corre-
sponding van der Waals force is repulsive for the silicarwaterrair triple layer. For an
unstable film, the thin film must drain then rupture and the resulting three phase line of
contact ŽTPLC, vapour–water–solid. must expand to form a wetting perimeter before
the particle can adhere to the bubble.
Each of these events will have a characteristic time associated with them, the sum of
which must be less than the contact time Ž t c . between the bubble and the particle if
flotation is to occur. The contact time is generally of the order of 10y2 s or less.
J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164 137

The induction time, t ind is normally taken as the time required for bubble–particle
attachment to occur, once the two are brought into contact, i.e., it is the sum of the thin
film drainage and TPLC spreading times Ž t film and t TPLC . and is synonymous with the
attachment time. Rupture is a very fast process and is not a significant contributor to
t ind .
2.2. Film drainage
When a bubble is pressed against a solid surface, through water, the intervening film
is generally not plane parallel. Rather the edge of the film thins quickly and a small,
thicker dimple is trapped in the centre, for the bubble is deformable. This is essentially a
kinetic phenomenon, caused by the flow being greatest at the very edge of the film in
the initial stages of drainage. The existence of this dimple has been detected by
techniques such as laser interferometry. Hydrodynamic theories attempting to describe
the profile and evolution of the dimple have been proposed but with very limited success
in describing experimental data.
Most approaches have not considered the rigorous differential equation describing the
drainage rate ŽBuevich and Lipkina, 1978., but instead have used various approxima-
tions to derive simpler differential equations which are amenable to analytical solution.
Approximate models Žand the equations derived from them. are as follows:
Ža. Frankel and Mysels Ž1962. assumed that the rate of flow through the barrier ring
per unit length of periphery was independent of the radius. Using this assumption, they
derived equations for the film thickness, h o , at the centre of the draining film, and h b , at
the barrier ring, as a function of time, t:
1
0.0096h rc6 4
ho s
ž g rb t / 1
Ž 3.

0.0090h rc2 r b 2
hb s
ž gt / Ž 4.
where r b is the radius of the air bubble, rc is the radius of the barrier ring, and
measured values are expressed in cgs units. g is the surface tension of water and h is its
viscosity.
Unfortunately, this model predicts that the shape of the dimple becomes more
pronounced as thinning progresses, which is the reverse of the experimental observation
ŽHewitt et al., 1993..
Žb. Hartland and Robinson Ž1977. approximated the shape of the dimple by three
linked parabolas, which led to the following equations for the film thickness as a
function of time:
1
2ph rc6 4
h o s 0.117
ž /g rb t
1
Ž 5.

r bh rc2 2
h b s 0.370
ž /
2pg t
. Ž 6.
138 J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164

Žc. Dimitrov and Ivanov Ž1978. assumed that all the energy dissipation occurred in
the barrier ring, and from this derived a model which predicted:
1
r bh rc2 2
ho s
ž / gt
. Ž 7.
Žd. Jain and Ivanov Ž1980. developed a ‘step’ model for the draining film, which
predicted:
1
r bh rc2 2
h b s 0.433
ž / gt
. Ž 8.
Že. Chen and Slattery Ž1982., Lin and Slattery Ž1982. and Chen Ž1984. produced a
series of models in which they improved on Hartland’s approach by redefining the
boundary conditions to produce a fourth-order differential equation which is identical in
form with that published earlier ŽBuevich and Lipkina, 1978.. To solve the equation, Lin
and Slattery Ž1982. postulated that the shape of the dimple was independent of time at
some early stage in the drainage, and adjusted the arbitrary parameters in their solution
so that the predicted curve fitted an experimentally measured curve ŽPlatikanov, 1964. at
an early time.
Hewitt et al. Ž1993. have examined the drainage behaviour between a hydrophilic
quartz surface and an approaching air bubble, using scanning interferometry. In Fig. 2,
their experimental drainage rate curves in doubly distilled water in the centre of the

Fig. 2. Film thickness in the centre of the draining film in conductivity water against time after film formation
is shown for film radius values of ŽB. 0.22=10y3 m, ŽI. 0.24=10y3 m, Ž'. 0.26=10y3 m and Ž^.
0.32=10y3 m. ŽFrom Hewitt et al., 1993, with permission..
J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164 139

Fig. 3. The inverse of the square of the film thickness Žat the centre of the dimple. against time for film radius
values of ŽB. 0.22=10y3 m, ŽI. 0.24=10y3 m, Ž'. 0.26=10y3 m and Ž^. 0.32=10y3 m, i.e., for the
same data as shown in Fig. 2. ŽFrom Hewitt et al., 1993, with permission..

draining film are shown for a series of experiments with increasing film radius at a
hydrophilic quartz surface. For the early drainage rates Žfor times - 50 s., a plot of the
inverse of the square of the film thickness against time results in a straight line ŽFig. 3.,
the gradient of which increases with decreasing film radius. Absolute values of the
gradient in all cases exceed that predicted by either Reynolds theory ŽReynolds, 1886. or
the theory of Dimitrov and Ivanov Ž1978. as shown in Table 2.
Observation of the drainage pattern which evolves when the bubble was pressed
against the quartz plate did not reveal any detectable asymmetry in the bubble ‘shape’.
Moreover, scanning the draining film on either side of the centre did not give detectably

Table 2
The experimental gradient derived from the plot of Žfilm thickness.y2 vs. time is shown for various film radii
observed; theoretical calculations of the gradient are shown for the theories of Reynolds Ž1886. Žtheoretical
slope 1. and Dimitrov and Ivanov Ž1978. Žtheoretical slope 2.
Film radius Experimental slope Theoretical slope 1 Theoretical slope 2
10y3 m 10 10 my2 sy1 10 13 my2 sy1 10 12 my2 sy1
0.22 2.19 1.38 1.30
0.24 1.93 1.16 1.09
0.26 1.53 0.99 0.93
0.32 1.18 0.65 0.61
140 J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164

different drainage rates at the same radii with all other conditions held constant. The two
surfaces Ži.e., airrwater and quartzrwater. clearly ‘sense’ one another at large distances
of separation, exhibit linear inverse square behaviour of film thickness against time, but

Fig. 4. Film thickness of the drainage film shown against time after film formation is shown for drainage at the
centre Ža. and the boundary ring Žb. at a hydrophilic surface in aqueous sodium chloride solutions of
concentrations of ŽB. 10y3 mol dmy3 , ŽI. 5=10y4 mol dmy3 , Ž'. 2.5=10y4 mol dmy3 and Ž^. 10y5
mol dmy3 sodium chloride solutions. ŽFrom Hewitt et al., 1993, with permission..
J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164 141

yield gradients which differ from extant theoretical predictions, even in the case of film
drainage for high-purity water.
Drainage rates at both the film centre Ž r s 0. and at the boundary ring Ž r s 0.20 =
10y3 m. at the quartz surface, as a function of salt concentrations, are shown in Fig. 4.
The drainage rate at the centre of the film decreases with increasing salt concentration
whereas at the boundary ring the reverse is true.
Quantitatively, this effect can be explained in terms of the decreased electrostatic
repulsion with increasing salt concentration, allowing the film to drain more rapidly at
the boundary ring. The drainage rate at the centre is then controlled by the rate at which
the aqueous solution can pass through the boundary ring. Film profiles for 2.5 = 10y4
mol dmy3 and 10y3 mol dmy3 salt solutions are shown in Fig. 5. These profiles
provide a graphic illustration of the point at which the ‘dimple’ becomes more
pronounced as the salt concentration is increased.
In Fig. 4b, particularly, it is clear that the influence of salt on drainage rate is sensed
at large distances of separation Ži.e., a micrometer or more.. This is also evident in Fig.
4a, albeit not as graphically. Plots of inverse square of film thickness against time are
given in Fig. 6a and b for drainage at the centre and the boundary ring, respectively. The
fit at the centre becomes more linear as the salt concentration increases, but still diverges
significantly from theory. These models predict 1rt n dependence with n s 0.25 at the
centre and n s 0.5 at the boundary ring, in disagreement with experiment. At equilib-
rium, the bubble surface flattens and, as noted by others ŽRead and Kitchener, 1969;
Blake and Kitchener, 1972., the system may be described as an asymmetric interface of
two parallel plates Žair and quartz., interacting through a thin liquid film.

Fig. 5. Film profiles are shown for drainage at a hydrophilic surface as a function of film radius shown for
2.5=10y4 mol dmy3 Žopen symbols, solid lines. and 10y3 mol dmy3 Žclosed symbols, dashed lines. sodium
chloride solutions for times of ŽI. 20, Ž`. 30, Ž^. 50, Ž`. 100 and ŽI. 200 s after film formation. ŽFrom
Hewitt et al., 1993, with permission..
142 J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164

Reasonable agreement is obtained between experimental equilibrium film thicknesses


Ži.e., when drainage has ceased. and those calculated using the DLVO theory. Compres-
sion of the double layer qualitatively explains the trends in the observed changes in the
drainage rates with variations in the ionic strength of the aqueous solution, however, the
J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164 143

distances at which the differences in drainage rates become apparent are an order of
magnitude greater than could be predicted or explained using the static or equilibrium
DLVO model. Dissolved gas bubbles trapped on the solid surface and roughness may be
implicated.
Surface deformation of bubble surfaces can also occur under the influence of
electrostatic interactions Žand possibly other surface forces as well., aside from any
kinetic effects, as shown by Miklavcic et al. Ž1995..
Experimental evidence relating to film drainage in systems where soluble surfactants
are present is rather equivocal. Adsorption and desorption processes coupled with
possible molecular reorientation make any theoretical interpretation difficult. Unfortu-
nately, these are the very systems which are of primary interest to mineral processing.
Furthermore, additional complications ensue when one considers a particle approaching
a bubble in flotation.
The nature of the bubble surface Ži.e., whether it is mobile or immobile. will
influence the evolution of the thin film between bubble and particle. This makes any
solution of the equation for film drainage difficult, particularly in the case of the angular
particles that are normally present in flotation. It is worth recalling at this point the
observations that smooth spheres float more slowly than angular particles under other-
wise identical conditions, presumably because the asperities on the angular particles lead
to increased film drainage rates andror rupture ŽBlake and Ralston, 1985; Anfruns and
Kitchener, 1986..
An unstable film arises when there is a net attractive force between the particle and
the bubble. This normally occurs when there is an attractive hydrophobic force involved,
for the van der Waals and electrostatic forces Žthe ‘DLVO’ components. are repulsive,
except in rare circumstances. The measurement of this hydrophobic force and its
theoretical origins are subjects of intense research effort. In recent times, it has become
possible to measure the hydrophobic force, in a configuration relevant to the flotation
process, by attaching a small particle to the cantilever in an atomic force microscope
ŽAFM. ŽFig. 7.. The particle is then pressed against a captive bubble and the force–sep-
aration distance profile determined, for example, as shown by Butt, 1994, Ducker et al.,
1994 and Fielden et al. Ž1996.. In this fashion, the various surface forces may be
explored.

2.3. Forces between bubble and particle

The interaction between a silica particle and an air bubble is shown in Figs. 8 and 9
ŽFielden et al., 1996.. It is monotonically repulsive on approach, in agreement with

Fig. 6. The inverse of the square of the film thickness against time after film formation is shown for drainage
at the centre Ža. and at the boundary ring Žb. at a hydrophilic surface in aqueous sodium chloride solutions of
concentrations of ŽB. 10y3 mol dmy3 . ŽI. 5=10y4 mol dmy3 , Ž'. 2.5=10y4 mol dmy3 and Ž^. 10y5
mol dmy3 sodium chloride solutions. Also shown are the models of Ž ____ . Frankel and Mysels Ž1962. and
Ž- - - -. Hartland and Robinson Ž1977. for drainage at the centre Ža.. In the case of drainage at the boundary ring
Žb., models are shown for the Ž ____ . Frankel and Mysels Ž1962. model and the Ž- - - -. Jain and Ivanov Ž1980.
model Žsee text.. ŽFrom Hewitt et al., 1993, with permission..
144 J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164

Fig. 7. Schematic diagram of experimental arrangement for the measurement of forces between a particle and a
bubble with an atomic force microscope. The bubble and particle diameters were approximately 600 and 3
mm, respectively, and the cantilever typically had a spring constant of 0.050 Nrm. ŽFrom Fielden et al., 1996,
with permission..

earlier bubble against plate observations ŽRead and Kitchener, 1969; Blake and Kitch-
ener, 1972.. The repulsion is electrical in origin, revealed through the influence of

Fig. 8. Raw data for the air bubble–silica interaction in water. The interaction is monotonically repulsive on
approach while on separation there is clear evidence of adhesion with Fdet s 3 mNrm Žnormalised force.. The
loading force, Fl , prior to separation was 6 mNrm. B denotes the region of particle—‘constant compliance’
region where the loading force is being monotonically increased. A is the region where repulsive forces are
evident. C relates to the bubble–particle adhesion which becomes evident when the bubble is moved away
from the particle. ŽFrom Fielden et al., 1996, with permission..
J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164 145

Fig. 9. Normalised force Ž Fr R . vs. separation Ž D . for air bubble–silica interaction in aqueous electrolyte.
Symbols represent experimental data measured in 10y4 M Ž', ^. and 10y2 M Ž`. sodium chloride. The
lines correspond to constant charge Ž . and constant potential Ž- - -. fits to DLVO theory. The fitting was
performed with the asymmetric DLVO program of Grabbe and Horn Ž1993.. cp was obtained from force
measurements between silica particles under identical conditions and was y100 and y27 mV at sodium
chloride concentrations of 10y4 and 10y2 M, respectively, with c b maintained at y34 mV. The analytical
concentration of electrolyte was used in the fitting procedure. The Hamaker constant for the interaction was
y1.0=10y2 0 J. In the inset, the detachment force–loading force data, normalised by particle radius, R p ,
measured on separation of the air bubble and particle are displayed. ŽFrom Fielden et al., 1996, with
permission..

electrolyte concentration. For an air bubble and silica, upon removal, a small adhesion
Ž- 4 mNrm normalised force. was typically observed which was found to be dependent
on the loading force, Fl , prior to removal, and on the electrolyte concentration.
The contrast between the approach and removal curves with respect to the sign of the
interaction is significant. An attractive ‘jump’ may occur on approach but may not be
detected in the experiment. A monotonic repulsion implies the presence of a stable thin
film, and an attractive force would be required for the removal of this aqueous film. An
adhesion is generally observed upon removal and it is possible that the reversal of the
piezoscanner, and attached bubble, perturbs the film and initiates drainage. The magni-
tude of the force ŽFig. 9. is not well described by theory. While there is some
146 J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164

uncertainty in the values of bubble potential used to generate the theoretical fits, the
greatest uncertainty relates to the difficulties inherent in deconvoluting the AFM data.
Although the ‘constant compliance’ slope is typically found to be acceptably linear ŽFig.
8., given the deformability of the bubble interface, its significance is not as clear as for
the interaction of nondeformable surfaces. The point of deviation of the approach curve
from the constant compliance slope cannot simply be equated to zero separation. For a
monotonically repulsive force prior to constant compliance, the onset of constant
compliance will correspond to some finite but indeterminate separation. Note that the
scale of this separation is significantly less than the range of the operative surface
forces; that is, the film is thin rather than thick. This underestimation of the separation
means that the measured force is less than predicted by theory. Conversely, the
deformation of the bubble surface due to an electrical repulsion ŽMiklavcic et al., 1995.
will lead to flattening, an increase in the area of interaction, and consequently, an
increase in the measured force. In the classic large bubble against plate measurements,
the system may be described as an asymmetric interface of two parallel plates interact-
ing through a liquid film ŽBlake and Kitchener, 1972; Hewitt et al., 1993.. Bubbles of
the smallest convenient diameter were used by Fielden et al. Ž1996. because their higher
Laplace pressure minimised the degree of deformation. Since the position of the
liquidrvapour interface was not directly measured, it was also likely that the deformabil-
ity of the bubble was not correctly accounted for in the deconvolution procedure. In this
context, the reasonable agreement between experiment and theory ŽFig. 9. is probably
due to the mutually compensating nature of these unknown factors.
When comparing experiment and theory ŽFig. 9., it is tempting to infer that on
approach both the bubble and silica particles are interacting under constant charge
conditions. However, if this were the case, then one might have expected the experimen-
tal data to increase sharply at small separations, where the flattening of the bubble would
be greatest ŽMiklavcic et al., 1995.. This is clearly not the case, with the experimental
data maintaining exponentiality over the entire range of separation. Intuitively, one
would expect the bubble and particle to interact at constant potential and constant
charge, respectively.
There are two possible explanations for the small adhesion observed between an air
bubble and silica. If one, or both, of the surfaces were interacting at constant potential,
then DLVO theory predicts that a charge reversal, for surfaces with potentials of
identical sign but different magnitude, should occur as the separation is decreased, on
the surface of lower potential. Given that the van der Waals force is monotonically
repulsive the resulting interaction would have a primary minimum, and a stable air
bubble–particle aggregate would be formed with a thin Ž3.5 nm. intervening liquid film.
This is the so-called ‘contactless flotation’ of Derjaguin and Dukhin Ž1960. and
Derjaguin et al. Ž1984. ŽFig. 10.. The resulting adhesion would be expected to decrease
with ionic strength and to be independent of loading force. While the measured
detachment force ŽFig. 9. was found to broadly correlate in this way with ionic strength,
a clear dependence on loading force was observed. If the silica surface is not hy-
drophilic, a possibility discussed by Ducker et al. Ž1994., and the receding water contact
angle is greater than zero, then the role of capillary forces must also be considered if
surface force, kinetic, and hydrodynamic factors enable establishment of a three phase
J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164 147

Fig. 10. DLVO calculation for the interaction of a silica particle and an air bubble in 10y4 M univalent
electrolyte. The potential of the silica particle was y100 mV, and that of the air bubble, y34 mV at infinite
separation. The charge on silica and the potential of the air bubble were held constant as the separation was
decreased. The Hamaker constant for the interaction was y1.0=10y2 0 J. ŽFrom Fielden et al., 1996, with
permission..

line ŽTPL.. Once a TPL is formed, capillary forces will dominate the adhesion ŽFig. 11.
as the contact angle increases, particularly because of the effect of contact angle
hysteresis. If a TPL has receded across a surface, either due to spontaneous dewetting or
forced movement due to an externally applied load, on separation, the contact line will
remain pinned ŽFig. 12. until the TPL subtends the advancing contact angle. The
expected adhesion, which may be readily calculated Žsee below under detachment. will
depend on loading force as well as on electrolyte concentration Žsince the contact angles
are affected, Fokkink and Ralston, 1989.. For the salt concentration range examined
here, the changes in contact angle are expected to be small ŽFokkink and Ralston, 1989.
Ža few degrees.. For the silica–bubble interaction, the detachment force–loading force
data may indeed be reasonably fitted by the capillary force given an advancing contact
angle of 108 and a finite receding angle close to 08. The possibility of contamination
cannot be ruled out. One of the weaknesses of the colloid probe technique is that
because of the small surface area of the particle, and additionally, in the present case the
bubble, it is almost impossible for surfaces to be kept entirely contaminant free. Given
that silica surfaces are well known to contaminate very readily, as observed from contact
angle studies, it is entirely possible that in this work, slight contamination would give
rise to an advancing contact angle of 108, providing support for a capillary force
148 J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164

Fig. 11. Variation of capillary force, Fcap , with v 8 Žparticle location. and with water contact angle, u in air
bubble–particle interaction. Fcap r R p s 2pg lv sinyŽ180y v .sinŽ180y v q u .. ŽFrom Fielden et al., 1996,
with permission..

dominated adhesion. It would be convenient if the adhesion could be attributed to either


surface or capillary forces simply on the basis of the magnitude of the measured
adhesion. However, the magnitudes of the expected forces are very similar for the two
cases.

2.3.1. Dehydroxylated silica


The interaction between dehydroxylated silica and an air bubble ŽFig. 13. was
repulsive at large separations but became attractive as the separation decreased, and a
significant adhesion was measured on removal. The force curves obtained for this
interaction were found to be quite reproducible with respect to successive approach—re-
moval cycles. A dependence of the interaction on electrolyte concentration was observed
while the detachment force was an order of magnitude greater than that for the
undehydroxylated silica–air bubble interaction. The detachment force was found to
depend on loading force, but there was no clear dependence of adhesion on electrolyte
concentration evident. DLVO theory does not predict the dehydroxylation of silica to
cause a significant change to the interaction, as only the potential of the particle surface
Žobtained here from the nondeformable solid interaction. has reduced slightly. However,
there is clearly a marked change in the measured interaction on approach. If dehydrox-
ylation of the silica surface caused it to interact under constant potential conditions, then
charge reversal of the bubble surface could occur at longer range and increase the depth
J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164 149

Fig. 12. Loading–unloading cycle for contact angles of 0 and 328, as a function of v 8 corresponding to the
values of the receding and advancing contact angles, respectively, of water on dehydroxylated silica measured
in this work. Fl and Fdet denote the loading force and detachment force, respectively. ŽFrom Fielden et al.,
1996, with permission..

of the primary minimum. However, this is highly unlikely and, in the case of dehydrox-
ylated silica, the attraction at decreasing separation is due to a hydrophobic force. The
uncertainties in the experimental data regarding deconvolution did not permit a quantita-
tive analysis of this attractive component. An electrical force cannot be used to explain
the detachment force for dehydroxylated silica, simply because of the magnitude of the
measured adhesion, it is likely that capillary forces are responsible. If the hydrophobic
attraction dominates the van der Waals repulsion at short range, then the aqueous film
separating the particle and bubble would be expected to thin rapidly and a TPL would be
established. Subsequently, capillary forces are expected to dominate the interaction. If
instead the van der Waals repulsion were to dominate the hydrophobic attraction, then a
primary minimum would occur and a stable thin film would exist between the
bubble–particle aggregate. For a stable hydrophobic attraction, this latter possibility
would give rise to a single-valued, electrolyte independent detachment force, in sharp
contrast to that observed. This explanation is also at odds with recent studies probing the
nature of hydrophobic surfaces ŽChristenson and Claesson, 1988; Wood and Sharma,
1995.. However, because hydrophobic attractions are not yet completely understood and
are found to be variable in magnitude and range of operation even within a single
experiment, this possibility cannot be entirely dismissed. In contrast, the alternative
scenario supporting the formation of a TPL is given strong support because there is
quantitative agreement between the measured and calculated values of the detachment
150 J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164

Fig. 13. Normalised force Ž Fr R . vs. separation Ž D . for interaction of dehydroxylated silica with an air bubble
in aqueous electrolyte. Symbols represent experimental data measured in water Ž = ., 10y4 M Ž', D . and 10y2
M Ž`, v . sodium chloride. The lines correspond to constant charge Ž . and constant potential Ž- - -. fits to
DLVO theory. cp was obtained from force measurements between dehydroxylated silica particles under
identical conditions and was y85 and y32 mV at sodium chloride concentrations of 10y4 and 10y2 M,
respectively, with c b maintained at y34 mV. The analytical concentration of electrolyte was used in the
fitting procedure. The Hamaker constant for the interaction was y1.0=10y2 0 J. In the case of 10y2 M
electrolyte, an outward shift of 15 nm for the fitted data allowed coincidence with the experimental data. This
shift provides extra evidence of TPL formation. In the inset, the detachment force–loading force data
measured on separation of the air bubble and particle is displayed. In this case, the dashed line corresponds to
the capillary force calculated for contact angles of 08 Žreceding. and 298 Žadvancing.. The corresponding
values of the contact angles, measured by the dynamic Wilhelmy method, were 08 Žreceding. and 328
Žadvancing.. ŽFrom Fielden et al., 1996, with permission..

force-loading force relationship ŽFig. 13 inset.. This was obtained assuming an advanc-
ing contact angle of 298 compared with the measured value of 328 ŽFig. 12..

2.4. TPLC

The kinetics of movement of the three phase line of contact ŽTPLC. are of central
importance in many processes, apart from flotation. During the movement of the TPLC,
a dynamic contact angle is established. Irrespective of whether the ‘surface chemical’,
J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164 151

‘hydrodynamic’ or mixed ‘surface chemicalrhydrodynamic’ approaches are used, there


is as yet no general theory which adequately describes TPLC kinetics on all surfaces
ŽBlake, 1993; Hayes and Ralston, 1993.. One cannot generally calculate ab initio what
the spreading velocity of the TPLC will be when an air bubble spreads over a mineral
surface immersed in water in the presence of a surfactant. Part of the problem at least is
due to the fact that poorly characterised experimental systems have been used, where
any generalisation has been obscured by the same time-dependent adsorptionrdesorp-
tionrmolecular reorientation processes which complicate thin film drainage rate studies.
Physical and chemical surface heterogeneities on the particle surface also strongly
influence the TPLC kinetics. Certainly, experimental evidence for t TPLC indicates that it
makes a significant contribution ŽSchulze and Stechemesser, 1997..
At present, only the crudest estimates of film drainage time, t film , and t TPLC can be
made. Hence, various experimental methods for determining t ind are frequently resorted
to. Ye and Miller have developed a potentially valuable approach to the calculation of
contact times, based on bubble deformation and restoration and have linked this to
induction time measurements and the flotation of coal ŽYe and Miller, 1989..
Experimental methods for determining induction times are generally based on either
pressing a bubble against a smooth mineral surface or against a bed of particles. The
disadvantages of all current methods for determining t ind include: Ž1. insufficient
understanding of the process of bubble deformation and energy dissipation during
bubblerparticle collision; Ž2. insufficient information concerning the behaviour of the
attractive hydrophobic forces during the bubblerparticle interaction Že.g., how the thin
film of liquid evolves with time, during the time a particle slides or rolls around a
bubble; it may well be incorrect to assume that bubble–particle interaction ceases when
the particle passes the bubble equator.; Ž3. data on t film , e.g., influence of surfactant type
and concentration on thin-film drainage mechanisms and rate; and Ž4. data on t TPLC as a
function of hydrophobicity, physical and chemical surface heterogeneities and surfactant
type. Recent work by Nguyen et al. Ž1997. indicates that a pseudo-line tension may play
a role in expansion of the three phase contact line.
The most appropriate method for determining induction times is probably through
direct observation of bubblerparticle interactions in a flotation cell under well-defined
conditions ŽSchulze, 1984.. The necessary theory can then be developed. For the
present, the Sutherland and similar approaches serve as useful approximations in
determining t ind from experimental flotation data of the type normally generated.
Kinetic effects certainly have a strong influence on bubble–particle collision and
attachment efficiencies. The latter provide the key to selective separations in flotation.
We now focus on the experimental determination of attachment efficiencies and discuss
how they may be described theoretically.
2.5. Attachment efficiency
2.5.1. Experimental measurement
When flotation experiments are conducted within the flotation domain where Es s 1
ŽHewitt et al., 1995; Dai et al., 1998., as described below, then the attachment efficiency
Ž Ea . is the ratio of the collection efficiency Ž E . to the collision efficiency Ž Ec ., i.e.,
Ea s ErEc , where E can be experimentally measured. Ec was calculated using the
152 J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164

Generalised Sutherland Equation ŽDai et al., 1998.. This enables one to obtain experi-
mental attachment efficiencies accurately and thus to test attachment efficiency models
with confidence.

Fig. 14. Calculated ŽEq. Ž21.; curves. and experimental Žsymbols. attachment efficiency as a function of
particle size for particles with various advancing water contact angles at different ionic strengths. d b s1.52
mm, pH s 5.6. ŽFrom Dai et al., 1998, with permission..
J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164 153

Experimental attachment efficiency results determined as described above under


various conditions are now presented ŽDai et al., 1998.. These results are then fitted to
the modified Dobby–Finch ŽDobby and Finch, 1987. and the Yoon–Luttrell ŽYoon and
Luttrell, 1989. attachment efficiency models and the outcomes discussed.
A single bubble flotation experiment technique has been used for the determination of
the collection efficiency. The samples used in the experiments were quartz particles
methylated to different but known advancing water contact angles. The details concern-
ing the experiments have been described elsewhere ŽHewitt et al., 1995; Dai et al.,
1998..
Fig. 14 indicates that for a given particle size, the attachment efficiency monotoni-
cally increases with contact angle, as well as with decreasing particle size, if the contact
angle is kept constant. The influence of particle surface hydrophobicity on attachment
efficiency can be explained by its influence on the interaction energy between a particle
and a bubble since the attractive hydrophobic interaction increases as the particle surface
hydrophobicity increases. This is accompanied by a decrease in induction time.
As is seen in Figs. 14 and 15, the attachment efficiency increases with electrolyte
concentration. An increase in electrolyte concentration reduces the interaction potential
energy between the particle and bubble, as well as the critical film rupture thickness and
increases the film drainage rate.
The increase in attachment efficiency with electrolyte concentration is more pro-
nounced for particles with greater hydrophobicity. Furthermore, as particle surface
hydrophobicity and electrolyte concentration are increased, the attachment efficiency
becomes less dependent upon particle size. In particular, for particles with contact angles

Fig. 15. Calculated ŽEq. Ž21.; curves. and experimental Žsymbols. attachment efficiency as a function of
particle size for particles with various advancing water contact angles Ž ua . at different ionic strengths and
bubble sizes. Curve 1 and symbol (: ua s 428, w l x s 0, d b s1.00 mm; curve 2 and symbol v: ua s 428,
w l x s 0, d b s 0.77 mm; curve 3 and symbol I: ua s 468, w l x s 0, d b s 0.77 mm; curve 4 and symbol B:
ua s 748, w l x s 0, d b s 0.77 mm; curve 5 and symbol ^: ua s 748, w l x s10y2 M, d b s 0.77 mm; pH s 5.6.
ŽFrom Dai et al., 1998, with permission..
154 J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164

larger than 708 and at wKClx s 10y2 M or higher, the attachment efficiency appears to be
independent of particle size.
The influence of bubble size on attachment efficiency is shown in Fig. 15. Fig. 15
indicates that under otherwise similar conditions, smaller bubbles always give higher
attachment efficiency than do larger bubbles. This is in good agreement with the
literature. For example, Yoon and Luttrell Ž1989. reported that for bubbles with diameter
larger than about 0.35 mm, the attachment efficiency decreases with increasing bubble
size. Hewitt et al. Ž1995. showed that bubbles of 0.75 mm in diameter always give
higher attachment efficiency than do 2.00 mm bubbles. In addition, Fig. 15 shows that
the increase in attachment efficiencies caused by decreasing bubble size is more
pronounced for smaller particles than for larger particles. Furthermore, Fig. 14 indicates
that as the particle contact angle increases, the difference in the attachment efficiency
between d b s 0.77 mm and d b s 1.52 mm becomes smaller. The attachment efficiency
is less dependent upon bubble size for strongly hydrophobic particles, in agreement with
the results of Hewitt et al. Ž1995..
Figs. 14 and 15 show that for strongly hydrophobic particles, the influence of bubble
size on attachment efficiency becomes weaker when the electrolyte concentration is
increased. In fact, for the sample with a contact angle of 748 and at a KCl concentration
of 10y2 M, the difference in attachment efficiency between the two bubble sizes is
within experimental error and the value of the attachment efficiency equals unity.
2.5.2. Attachment efficiency models
The Dobby–Finch attachment efficiency model ŽDobby and Finch, 1987. is based on
the relative magnitude of the induction time and the sliding time. The latter is the time
taken for the particle to slide around the bubble surface, until it moves away from the
bubble surface. Bubble surface deformation and thus particle rebound are ignored for
fine particles. Assuming that particle–bubble collision occurs evenly over the section of
the bubble surface between u s 0 and u s uc where u is measured from the front
stagnation point of the bubble, or the north pole in the case of a rising bubble, and uc is
the maximum possible collision angle, these authors proposed that the attachment
efficiency Ž Ea . is the ratio of the area corresponding to ua to the area corresponding to
uc . The basic equation of the model is expressed as
sin2ua
Ea s 2 Ž 9.
sin uc
where ua , termed the adhesion angle is a specific collision angle where if a particle
collides at this angle, its sliding time is just equal to the induction time. So the angle ua
relates the sliding time and induction time to the attachment efficiency.
According to Dobby and Finch, the maximum collision angle Ž uc . is a function of
only the bubble Reynolds number. Expressions for uc have been derived by fitting the
experimental results of Jowett Ž1980. where data show that the value of uc decreases
linearly with the logarithm of the bubble Reynolds number. One obtains:
uc s 78.1 y 7.37 log R eb for 20 - R eb - 400; Ž 10 .
uc s 85.5 y 12.49 log R eb for 1 - R eb - 20; Ž 11 .
uc s 85.0 y 2.50 log R eb for 0.1 - R eb - 1. Ž 12 .
J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164 155

Nguyen Ž1994. improved this approach, accounting for the physical characteristics of
the particles in addition to the Reynolds number.
In this study, to be consistent with the assumptions made in the collision model ŽDai
et al., 1998., we used a maximum collision angle which is different to that used by
Dobby and Finch Ž1987.. This maximum collision angle, denoted by u t , is a complex
function of the bubble Reynolds number and satisfies the equation ŽDai et al., 1998..
1r2
sin2u t s 2 b Ž1 q b 2 . yb Ž 13 .
in which the dimensionless number b , after modification ŽDai et al., 1998. is defined as
4 d p rp
bs Ž 14 .
3 Kd b Ž rp y r f .

where d p and d b are particle and bubble diameters, rp and r f are particle and fluid
densities, respectively, K is the Stokes number which characterises the ratio of the
inertial forces of the particle to the viscous resistance of the fluid and is

rp P Õ b P d p2
Ks Ž 15 .
9h d b

in which Õ b is the bubble rising velocity and h is the dynamic viscosity of the fluid. Eq.
Ž14. can be rewritten as

12 d b rf 1
bs P P Ž 16 .
dp rp y r f R eb

where the bubble Reynolds number Ž R eb . is

R eb s Õ b d b r frh . Ž 17 .
As mentioned above, if a particle collides on the bubble surface at a certain angle and
then slides along the surface to the equator of the bubble, and if the sliding time is just
equal to the induction time, this particular collision angle is the adhesion angle Ž ua .. The
sliding velocity, or sliding time, of a particle depends on the fluid flow regime at the
bubble surface. It has been shown ŽDai et al., 1998. that the fluid flow at the bubble
surfaces under flotation conditions can be best described by potential flow. Therefore,
the sliding time model for potential flow conditions should be used for the determination
of the adhesion angle Ž ua ..
Dobby and Finch Ž1986. derived an equation for the sliding time Ž tsl . under potential
flow conditions.
dp q db u
tsl s
db
3 ž /
P ln cot
2
. Ž 18 .
2 Ž Õp q Õ b . q Ž Õ p q Õ b .
ž dp q db /
156 J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164

If the sliding time is substituted for the induction time Ž t ind ., the calculated angle is
then the adhesion angle. Hence, ua is determined by the equation
3
db

ua s 2 arccot exp t ind


2 Ž Õp q Õ b . q Ž Õp q Õ b .
ž dp q db / Ž 19 .
dp q db

3
db

ua s 2 arctan exp ytind


2 Ž Õp q Õ b . q Ž Õ p q Õ b .
ž dp q db / . Ž 20 .
dp q db

Substitution of Eqs. Ž13. and Ž20. into Eq. Ž9. yields

° db
3

~
sin2 2 arctan exp yA P d pB P
2 Ž Õp q Õ b . q Ž Õp q Õ b .
ž dp q db / •
dp q db
¢ ß
Ea s 1 Ž 21 .
2
2 b Ž1 q b 2 . y b

in which b is determined by Eq. Ž16. and the induction time can be represented as

t ind s A P d pB Ž 22 .

in accordance with the well-documented dependence of the induction time upon particle
size Že.g., Schulze, 1984..
In the fitting process, the parameters A and B for the induction time are adjusted
independently so as to obtain best agreement between calculated and experimental Ea
values for each experimental condition. The values of A and B thus obtained are then
used to calculate the induction time. Finally, the calculated induction time and its
variation with various factors are compared with those reported in literature. Whether
agreement with literature is obtained or not is a critical assessment of the validity of the
modified Ea model ŽEq. Ž21...
The values of A and B may be determined through an iterative procedure as
described elsewhere ŽDai, 1998.. At the first stage, for a given experimental Ea vs.
particle size curve, an initial set of values of A and B was tried in Eq. Ž21. to produce a
calculated attachment efficiency Ž Eacal . since the value of b can be easily determined
J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164 157

from Eqs. Ž14. and Ž16.. Then Eacal was compared with the experimental attachment
efficiency Ž Eaexp .. If the difference between Eacal and Eaexp is larger than the preset
tolerable error, the values of A and B are independently changed and Eacal is calculated
again. This procedure is repeated until the difference between Eacal and Eaexp is smaller
than the tolerable error. The last set of A and B values is recorded and the same
procedure starts over for another Eaexp vs. d p curve. The aim of the above process is to
find out how the values of A and B Žand thus the induction time. change with process
parameters Žsuch as particle contact angle, bubble size and ionic strength..
The calculation using the above procedure shows that the value of B is almost
independent of particle contact angle, bubble size and ionic strength. This is in
agreement with the theory of Jowett Ž1980. and the experimental results of Ye and
Miller Ž1989.. Since these data are obtained by fitting to the experimental Ea which was
determined from experimental collection efficiency and a collision efficiency model, the
data are rather scattered due to experimental error and imperfection of the collision
model. However, as all values of B fall in the range between 0.45 and 0.75, they can be
approximately taken as a constant.
At the second stage, the computation program is such that a constant value of B is to
be sought for all Eaexp vs. d p curves and an individual value of A is to be determined for
each of these curves so as to get the best agreement between Eacal and Eaexp as a whole.
It turns out that a value of B s 0.6 is obtained. This value is very close to the average of
the values of B determined at the first stage and falls well in the range between 0 and
1.5, as theoretically proposed by Jowett Ž1980. but is smaller than the experimental
value of 1.8 to 2.2, as reported by Ye and Miller Ž1989..
The calculated attachment efficiencies based on a constant value of B Žs 0.6. and
individual values of A are displayed in Figs. 14 and 15 together with experimental
attachment efficiencies. In general, the agreement between calculated and experimental
values is satisfactory.
The Yoon–Luttrell ŽYoon and Luttrell, 1989. attachment efficiency model is also
based on the relative magnitude of the induction time and the sliding time, and thus, is
proposed for fine particles which do not rebound from the bubble surfaces. It has three
equations, each corresponding to a particular fluid flow regime. For the case of potential
flow conditions, the Yoon–Luttrell Ea model is described by the equation

y3Õ b t ind

Ea s
2
sin ua
s
½
sin2 2 arctan exp
dp q db 5 Ž 23 .
sin2uc sin2 90

y3Õ b t ind
where ua s 2 arctan exp
dp q db

is the expression for the adhesion angle ua . The ‘sin2 90’ in the denominator of Eq. Ž23.
implies the assumption that uc s 908, or in other words, particle–bubble collision occurs
uniformly over the entire upper half of the bubble surface. This has been shown to be
158 J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164

incorrect Že.g., Dai et al., 1998; Dobby and Finch, 1987.. If Eq. Ž13. is substituted for uc
Žor u t . and Eq. Ž22. for t ind , Eq. Ž23. becomes

y3Õ b A P d pB

Ea s
½
sin2 2 arctan exp
Ž dp q db .
1
5 . Ž 24 .
2
2 b Ž1 q b 2 . y b

A similar procedure to that used for the modified Dobby–Finch Ea model was
applied to Eq. Ž24. and the same calculation results as shown in the preceding section
were obtained. This means that the two models, after modification, are unified. As a
matter of fact, from Eqs. Ž9. – Ž12. and Ž23., it can be seen that the original concept of
the two models is very similar, the main difference being the value of the angle uc . This
difference has been eliminated by using the correct expression for the maximum
collision angle u t ŽEq. Ž3... Another difference is the expression for the angle of ua Žcf.
Eqs. Ž20. and Ž23... This difference originates from, in the Yoon–Luttrell model, the
neglect of the contribution of the particle settling velocity to its sliding velocity and the
identification of the latter to the tangential velocity of the fluid at the bubble surface. In
fact, if it is assumed that Õ b q Õp f Õ b and d b q d p f d b , Eqs. Ž20. and Ž21. reduce to
Eq. Ž23.. Since these approximations are valid only when d p - d b , such as in the present
study, very close results have been obtained from the two independent approaches. It is
obvious that the modified Dobby–Finch Ea model is more general than the modified
Yoon–Luttrell model. However, this does not mean that the modified Dobby–Finch
model can be applied to cases where coarse particles are involved since under such
circumstances, bubble surface deformation and thus particle rebound which are ignored
in the Dobby–Finch model may be important.
Neither of these models address the fundamental issues of thin film drainage, and
three phase contact line movement which are the major contributors to Ea , readily seen
through the link between t ind , t c Žor tsl ., t film and t TPLC
t ind Ž tfilm q t TPLC .
Ea s 1 y s1y Ž 25 .
tc tc
In arriving at a theoretical description of Ea , which correctly incorporates film
drainage and contact line movement, the angular dependence of the various forces must
be incorporated. This is yet to be attempted and is a major task for the future.

3. Stability and detachment

3.1. Flotation limits for coarse particles

The essential problem in understanding bubble–particle aggregate stability is to


determine whether or not the adhesive force, acting on the three phase contact line, is
J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164 159

large enough to prevent the destruction of the aggregate under the dynamic conditions
which exist in flotation.
It is important that the reader understands the physics of the problem before moving
on to a mathematical description. Let us consider a smooth spherical particle located at
the fluid interface. Once the equilibrium wetting perimeter has been established follow-
ing spreading of the three phase contact line, the static buoyancy of this volume of the
particle will act against the gravitational force ŽFig. 16.. The hydrostatic pressure of the
liquid column of height Z0 acts against the capillary pressure. The ‘other detaching
forces’ require further discussion—since they arise from the particle motion relative to
the bubble, velocity dependent drag forces will oppose the detachment of the particle
from the bubble. An analysis of these forces is extremely complex and has not been
reported to date.
The net adhesive force, Fad , is equal to the sum of the attachment forces, Fa , minus
the detachment forces, Fd , i.e.:
Fad s Fa y Fd . Ž 26 .
An equilibrium position is achieved if Fad is zero. The particle will not remain attached
to the bubble if Fad is negative but will be present in the liquid phase.
The mathematical description of the various forces which dictate the equilibrium
position of particles at liquid–vapour or liquid–liquid interfaces has followed an
evolutionary trail. Analogous processes of interest, for example, include pigment
‘flushing’, where a solid particle is induced to transfer from one liquid phase to another
by appropriate surface modification with surfactants and the stabilisation of emulsion
droplets by solid particles.
It was Princen Ž1969. who proposed the first extensive and generalised treatment of
the forces acting on a particle at fluid interfaces. This theory was developed further by
Schulze Ž1977, 1984..

Fig. 16. Location of a smooth spherical particle of radius, R p at a fluid interface. The parameters shown
enable the force and energy balance to be calculated Žsee text.. ŽFrom Schulze, 1984, with permission..
160 J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164

It is now pertinent to consider the case of a spherical particle at liquid–air interface.


The force balance will be used and then linked to the energy balance. We assume that
the system is in a quasistatic state and that the contact angle corresponds to that obtained
for a static system. The dynamic contact angle can depart significantly from the static
value, depending in part on the velocity of the TPLC. If the particle oscillates around its
equilibrium position, the TPLC would be expected to move to some extent.
Hence, a full analysis would need to account for the velocity dependent drag forces
mentioned above and link these to the contact angle dynamics. This is not a tractable
problem at present so that a simpler approach is necessary.
Let us suppose that a spherical particle of radius R p is attached to a bubble of radius
R b where R b is much greater than R p , as shown in Fig. 16. From understanding the
forces Že.g., capillary, buoyancy, capillary pressure, etc.. which operate on the particle,
it is possible to calculate the energy of detachment.
The energy of detachment, Edet , corresponds to the work done in forcing a particle to
move from its equilibrium position, h eq Ž v . at the liquid–vapour interface, as a function
of the central angle v ŽFig. 16., to some critical point, h crit Ž v ., where detachment
occurs and the particle moves into the liquid phase. The sum of the various forces, ÝF,
is related to Edet by:

Hhh Ž v .v ÝFd h Ž v . .
crit Ž .
Edet s Ž 27 .
eq

The detachment process takes place when the kinetic energy of the particle equals
Edet . The kinetic energy of the particle is given by Ž2r3.p R 3p rpVt 2 , where Vt is the
relative Žturbulent. velocity of the particle, acquired due to stresses on the bubble–par-
ticle aggregate in the turbulent field of the flotation cell, as the aggregate collides with
other bubbles or aggregates or due to other modes of excitation. Vt corresponds to the
velocity of gas bubbles in the flotation cell. rp and r fl refer to the densities of the
particle and fluid, respectively.
The maximum floatable particle diameter based on the kinetic theory, Dmax, K , is
given as:

3 h crit Ž v . 2 2 rp 3h
Dmax , K s 2
2prpVt 2 Hh
eq Ž v .
½ 3 ž
p R 3p r fl g 1 y
r fl
y cos 3v q
2 Rp
sin2v

1
3 2 2g 3
y
a 2 R 2p /
sin v sin Ž v q u . yp Ž R p sin v .
ž Rb
y 2 R b r fl g
/5 dh

Ž 28 .

where a is the capillary or Laplace constant,

pfl g
( g
.
J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164 161

Eq. Ž28. may also be solved by numerical integration or by plotting each of the
kinetic and detachment energies as a function of R p at constant g and rp and specified
Vt . This equation has been shown to adequately describe the detachment of a sphere
from a liquid–vapour interface by Schulze as well as the behaviour of hydrophobic
angular quartz particles, between approximately 35 to 120 mm in diameter, under
flotation conditions by Crawford and Ralston Ž1988..

3.2. Flotation limits for fine particles

The only theoretical study to date dealing with the limit of floatability of fine
particles was proposed by Scheludko et al. Ž1976.. The limit is the critical work of
expansion required to initiate a primary hold or three phase contact line during
bubble–particle approach, a requirement which is met by the kinetic energy of the
particles. The matching of these two quantities enables a minimum particle diameter,
Dmin, K for flotation to be obtained:
1
3k 2 3
Dmin , K s 2 Ž 29 .
Vt 2D rg  1 y cos u 4
where k is the line tension, opposing expansion of the three phase contact, D r is the
difference between particle and fluid densities and u is the contact angle at the TPLC. It
was Gibbs who first identified the importance of line tension, using as example the line
of intersection of the three surfaces of discontinuity which exist when two gas bubbles
adhere together.
Molecules which are present in a line have a free energy which is different from
those at a surface—in fact, there is an excess linear free energy and a linear tension in
an analogous fashion to that of excess surface free energy and surface tension.
In fact,
EF
ks ž /
EL T ,V ,W
Ž 30 .

where F is the Helmholtz free energy, L is the contact line, W is the thermodynamic
work, T is the temperature and V is the volume. The Young–Dupre equation becomes:
k
g S r V y g S r L s g L r V cos u " . Ž 31 .
r
The line tension is important for small contact radii and can oppose or reinforce
g L r V cosu. It counteracts the formation of the three phase contact line in Scheludko’s
theory, the latter neglecting thin film drainage and other hydrodynamic effects. The
experimental data of Crawford and Ralston Ž1988. for hydrophobic, angular quartz
particles between about 10 to 30 mm in diameter follow a general trend which is
predicted by Eq. Ž29.. Drelich and Miller Ž1992. have shown that acceptable agreement
is achieved if a pseudo-line tension, embracing surface heterogeneities, replaces k in
Eq. Ž29.. This in turn enables Dmin in Eq. Ž29. to be reexpressed in terms of a critical
bubble radius below which attachment does not occur. Reconciliation between theory
162 J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164

and experiment is then achieved, although the concept of pseudo-line tension needs to be
placed on a firmer experimental footing.
Hysteresis, the difference between static and dynamic contact angles caused by
surface heterogeneities, is important in both attachment and detachment. Contact line
pinning can occur during movement of the TPLC, inhibiting attachment or detachment,
i.e., the bubble–particle union may be stabilised. A quantitative description of the effect
of surface heterogeneities on TPLC movement is not yet available, although recent work
holds promise ŽPetrov et al., in press..

4. The future

In terms of our fundamental understanding, thin film drainage is poorly understood


when one of the interfaces is both physically and chemically heterogeneous, the other
deformable. The nature of the hydrophobic interaction between a particle and a bubble
requires both experimental and theoretical verification, particularly the influence of
dissolved gases. There is no reliable model at present to describe the movement of a
three phase contact line over a rough surface, nor to describe the thin film drainage
process between a bubble and a particle. The angle dependence of the interaction
between a bubble and particle must be addressed during film drainage. Thus, major
research challenges exist which, if they are to be successfully overcome, must embrace
systems where surfactants are both present and absent.
From a separation technology point of view, froth flotation will continue to be one of
the principal means by which ores are successfully beneficiated for many years to come.
Increasingly, the technique is also being used in the deinking of paper, soil remediation,
plastics recycling and heavy metal ion contamination, to name but a few examples. Both
research and practice are expected to accelerate strongly over the next decades as new
techniques and theoretical approaches are used.

References

Anfruns, J.F., Kitchener, J.A., 1986. Rate of capture of small particles in flotation. Trans. IMM Sect. C. 86,
C9.
Blake, T.D., 1993. In: Berg, J.C. ŽEd.., Wettability, Chap. 5. Marcel Dekker, New York.
Blake, T.D., Kitchener, J.A., 1972. Stability of aqueous films on hydrophobic methylated silica. J. Chem. Soc.
Faraday Trans. 1 Ž68., 1435.
Blake, P., Ralston, J., 1985. Particle size, surface coverage and flotation response. Colloids Surf. 16, 41.
Buevich, Y.A., Lipkina, E.Kh., 1978. Disruption of thin liquid films. Colloid J. USSR 40, 167.
Butt, H.J., 1994. A technique for measuring the force between a colloidal particle in water and a bubble. J.
Colloid Interface Sci. 166, 109.
Chen, J.D., 1984. Effects of London van der Waals and electrical double layer forces on the thinning of a
dimpled film between a small drop or bubble and a horizontal solid plane. J. Colloid Interface Sci. 98, 329.
Chen, J.D., Slattery, J.C., 1982. Effects of London van der Waals forces on the thinning of a dimpled film
between a small drop or bubble and a horizontal solid plane. AIChE J. 28, 955.
Christenson, H.K., Claesson, P.M., 1988. Cavitation and the interaction between macroscopic hydrophobic
surfaces. Science 239, 390.
J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164 163

Crawford, R., Ralston, J., 1988. The influence of particle size and contact angle in mineral flotation. Int. J.
Miner. Proc. 23, 1.
Crawford, R., Koopal, L.K., Ralston, J., 1987. Contact angles on particles and plates. Colloids Surf. 27, 57.
Dai, Z., 1998. Particle–Bubble Heterocoagulation, PhD Thesis, University of South Australia.
Dai, Z., Dukhin, S.S., Fornasiero, D., Ralston, J., 1998. The inertial hydrodynamic interaction of particles and
rising bubbles with mobile surfaces. J. Colloid Interface Sci. 197, 275.
Derjaguin, B.V., Dukhin, S.S., 1960. Theory of flotation of small and medium size particles. Trans. Inst.
Mining Met. 70, 221.
Derjaguin, B.V., Churaev, M.V., 1976. The definition of disjoining pressure and its importance in the
equilibrium and flow of thin films. Colloid J. USSR 38, 402.
Derjaguin, B.V., Dukhin, S.S., Rulyov N.N., 1984. In: Matijevic, E., Good, R.J. ŽEds.., Kinetic Theory of
Flotation of Small Particles. Surf. Colloid Sci., Vol. 13, p. 71.
Diggins, D., Fokkink, L.G.J., Ralston, J., 1990. The wetting of angular quartz particles. Colloids Surf. 44, 299.
Dimitrov, D.S., Ivanov, I.B., 1978. Hydrodynamics of thin liquid films. J. Colloid Interface Sci. 64, 97.
Dobby, G.S., Finch, J.A., 1986. A model of particle sliding time for flotation size bubbles. J. Colloid Interface
Sci. 109, 493.
Dobby, G.S., Finch, J.A., 1987. Particle size dependence in flotation derived from a fundamental model of the
capture process. Int. J. Miner. Proc. 21, 241.
Drelich, J., Miller, J.D., 1992. The effect of surface heterogeneity on pseudo-line tension and the flotation
limit of fine particles. Colloids Surf. 69, 35.
Ducker, W.A., Xu, Z., Israelachvili, J.N., 1994. Measurements of hydrophobic and DLVO forces in
bubble–surface interactions in aqueous solutions. Langmuir 10, 327924.
Fielden, M., Hayes, R., Ralston, J., 1996. Surface and capillary forces affecting air bubble–particle interac-
tions in aqueous electrolyte. Langmuir 12, 3721.
Fokkink, L.G.J., Ralston, J., 1989. Contact angles on charged substrates. Colloids Surf. 36, 69.
Frankel, S.P., Mysels, K.J., 1962. On the dimpling during the approach of two interfaces. J. Phys. Chem. 66,
190.
Grabbe, A., Horn, R.G., 1993. Double layer and hydration forces measured between silica sheets subjected to
various surface treatments. J. Colloid Interface Sci. 157, 375.
Hartland, S., Robinson, J., 1977. A model for the axisymmetric dimpled drawing film. J. Colloid Interface Sci.
60, 72.
Hayes, R.A., Ralston, J., 1993. Forced liquid movement on low energy surfaces. J. Colloid Interface Sci. 159,
429.
Hewitt, D., Fornasiero, D., Ralston, J., Fisher, L.R., 1993. Aqueous film drainage at the quartzrwaterrair
interface. J. Chem. Soc. Faraday Trans. 89, 817.
Hewitt, D., Fornasiero, D., Ralston, J., 1995. Bubble particle attachment. J. Chem. Soc. Faraday Trans. 91
Ž13., 1997.
Jain, R.K., Ivanov, I.B., 1980. Thinning and rupture of ring shaped films. J. Chem. Soc. Faraday Trans. 2 Ž76.,
250.
Jowett, A., 1980. Formation and disruption of bubble–particle aggregates in flotation in fine particles
processing. In: Somasundaran, P. ŽEd.., Proc. Int. Symp. Fine Particles Processing. AIME, AIMM New
York, Vol. 1, Chap. 37.
Lin, C.-Y., Slattery, J.C., 1982. Thinning of a liquid film as a small drop or bubble approaches a solid plane.
AIChE J. 28, 147.
Miklavcic, S.J., Horn, R.G., Bachmann, D.J., 1995. Colloidal interaction between a rigid solid and a fluid
drop. J. Phys. Chem. 99, 16357.
Nguyen, A.V., 1994. The collision between fine particles and single air bubbles in flotation. J. Colloid
Interface Sci. 162, 123.
Nguyen, A.V., Schulze, H.J., Ralston, J., 1997. Elementary steps in particle–bubble attachment. Int. J. Miner.
Proc. 51, 183.
Petrov, J., Ralston, J., Hayes, R.A., in press. Dewetting dynamics on heterogeneous surfaces. A molecular
kinetic treatment. Langmuir.
Platikanov, D., 1964. Experimental investigation on the dimpling of thin liquid films. J. Phys. Chem. 68, 3619.
Princen, A., 1969. In: Matijevic, E. ŽEd.., Surf. Colloid Sci., Wiley, New York, NY, Vol. 2, Chap. 2.
164 J. Ralston et al.r Int. J. Miner. Process. 56 (1999) 133–164

Read, A.D., Kitchener, J.A., 1969. Wetting films on silica. J. Colloid Interface Sci. 30, 391.
Reynolds, O., 1886. On the theory of lubrication. Philos. Trans. R. Soc. 177, 157.
Scheludko, A.D., 1967. Thin liquid films. Adv. Colloid Int. Sci. 1, 391.
Scheludko, A., Toshev, B.V., Bojadjiev, D.T., 1976. Attachment of particles to a liquid surface Žcapillary
theory of flotation.. J. Chem. Soc. Faraday Trans. 1 Ž72., 2815.
Schulze, H.J., 1977. New theoretical and experimental investigations on stability of bubble–particle aggregates
in flotation: a theory on the upper particle size of floatability. Int. J. Miner. Proc. 4, 241.
Schulze, H.J., 1984. Physicochemical Elementary Processes in Flotation, Elsevier, Amsterdam.
Schulze, H.J., Stechemesser, H., 1997 Žprivate communication..
Sutherland, K.L., 1948. Physical chemistry of flotation: XI. Kinetics of flotation process. J. Phys. Chem. 52,
394.
Wood, J., Sharma, R., 1995. How long is the long-range hydrophobic attraction?. Langmuir 11, 4797.
Ye, Y., Miller, J.D., 1989. The significance of bubble contact time during collision in the analysis of flotation
phenomena. Int. J. Miner. Proc. 25, 199.
Yoon, R.H., Luttrell, G.H., 1989. The effect of bubble size on fine particle flotation. Miner. Proc. Extract.
Met. Rev. 5, 101.

You might also like