Sunteți pe pagina 1din 18

The limits of ne particle otation

Tatu Miettinen
a
, John Ralston
b,
*
, Daniel Fornasiero
b
a
Outokumpu Research Oy, Kuparitie 10, P.O. Box 60, FIN-28101 Pori, Finland
b
Ian Wark Research Institute, University of South Australia, Mawson Lakes Campus, Mawson Lakes, Adelaide, SA 5095, Australia
a r t i c l e i n f o
Article history:
Received 21 September 2009
Accepted 14 December 2009
Available online 21 January 2010
Keywords:
Fine particle otation
Flotation
a b s t r a c t
Understanding the limits of ne particle otation is the key to the selective separation of ne mineral
particles. Fine particles have low collision efciencies with gas bubbles and oat slowly. There has been
a great deal of work aimed at overcoming the inefcient collision of small particles with rising air bub-
bles. This review deals with the inuence of bubble size, particle aggregation, different ow conditions,
particle induction time, as well as the action of surface and capillary forces on ne particlebubble cap-
ture. Recommendations for practice are given.
Crown Copyright 2009 Published by Elsevier Ltd. All rights reserved.
1. Introduction
1.1. Milestones in ne particle otation
In 1942, research conducted by Gaudin et al. demonstrated that
ne particles possess different otation properties from larger par-
ticles. Gaudin proposed that the rate of otation was independent
of particle diameter up to 4 lm, and proportional to the diameter
for particles in the range of 420 lm. However, this early observa-
tion was overtaken by research which concentrated upon the
development of collision models for larger particles.
The rst theoretical model to describe collision between a ne
particle and a bubble was developed by Sutherland (1948) in
1948. He proposed that the rate of otation could be found by
assuming that the ow around the bubble was the uniform motion
of an inviscid uid. Thus the streamlines around the bubble could
be calculated using potential ow theory. Assuming that particles
have no inuence on the oweld, bubbleparticle collision occurs
when the particle lies on a streamline which brings it within one
particle radius of the spherical bubble surface.
In 1961, Derjaguin and Dukhin (1961) produced a key paper on
ne particle otation dealing with the theory of otation of small
and medium-size particles. This paper included the effects of
hydrodynamics, surface forces and diffusiophoresis. Derjaguin
and Dukhin formally proposed that the overall rate of otation is
equal to the product of three probabilities or efciencies: bub-
bleparticle collision, attachment and stability.
Thirty years after Gaudins original proposition, Reay and Rat-
cliff (1973) suggested that two otation regimes might exist. They
proposed that the rst regime occurs for particles with diameters
greater than 3 lm, where the bubbleparticle collision efciency
increases with increasing particle size. For particles with a diame-
ter less than several microns, they become susceptible to Brownian
diffusion and enter a second otation regime. Brownian diffusion is
the main collision mechanism operative when particles approach
molecular dimensions. Diffusion is an effective mass transfer
mechanism, and the loss of inertia associated with ne particles
becomes a positive advantage in Brownian motion. Therefore,
one would expect the bubbleparticle collision efciency to in-
crease with a reduction in particle size in the diffusion regime
(Reay and Ratcliff, 1973).
In 1977, Anfruns and Kitchener published the rst measure-
ments of the collection efciency between ne particles and bub-
bles under potential ow conditions. This was the rst critical
test of Sutherlands collision theory under conditions where the
particle and bubble surface chemistry was controlled.
Recently, the data of Nguyen et al. (2006) showed that bubble
particle collection efciency is a minimum at a particle size around
100 nm or 0.1 lm. With larger particles, the interception and col-
lision mechanisms apparently dominate, while the diffusion and
colloidal forces control the collection of particles with a size smal-
ler than the detected transition size of 100 nm or so.
This short introduction is meant to sharpen the readers appe-
tite, for the problem of dealing with ne particles has puzzled min-
eral otation engineers for over a century. It is even more pressing
today, driven by energy and water minimization requirements, as
well as by mineralogy and liberation requirements of many ore
deposits.
2. Approaches for ne particle otation
The fundamental reason for the low otation rate of ne
particles is primarily due to their low collision efciency with
0892-6875/$ - see front matter Crown Copyright 2009 Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.mineng.2009.12.006
* Corresponding author. Tel.: +61 8 8302 3066; fax: +61 8 8302 3683.
E-mail address: John.ralston@unisa.edu.au (J. Ralston).
Minerals Engineering 23 (2010) 420437
Contents lists available at ScienceDirect
Minerals Engineering
j our nal homepage: www. el sevi er. com/ l ocat e/ mi neng
conventional otation bubbles of a given size and velocity (Trahar
and Warren, 1976; Fuerstenau, 1980; Yoon and Luttrell, 1989; Dai
et al., 2000). Several otation technologies have been developed,
which aim at increasing bubbleparticle collision efciency, either
by decreasing bubble size or by increasing apparent particle size.
2.1. Decreasing bubble size
There is a large body of experimental evidence showing that
bubbleparticle collection efciency (Anfruns and Kitchener,
1977; Yoon and Luttrell, 1986; Hewitt et al., 1995; Dai et al.,
1998, 1999) and otation rate (Bennett et al., 1958; Reay and Rat-
cliff, 1975; Ahmed and Jameson, 1985) increase with decreasing
bubble size. Unfortunately, the use of small bubbles also involves
disadvantages. Due to the low rising velocity of rather small bub-
bles with attached particles, very long otation times result and
hence require a substantial residence time in otation circuits. An-
other disadvantage is that the lifting force of small bubbles may be
too low to ensure process selectivity. Furthermore, it has been ob-
served that microbubbles cause high water recovery (Schwarz and
Grano, 2002), which increases the entrainment of gangue minerals
(Trahar and Warren, 1976; Liu and Wannas, 2004).
A decrease in bubble size can be obtained using different
methods, which can be divided into mechanical and physiochem-
ical approaches. Mechanical methods include the design of ota-
tion cells, i.e., the shape of the rotor and stator, and the gap size
between the rotor and stator, so that the gas bubbles produced at
the bottom of the otation cell can be dispersed into smaller bub-
ble sizes. A microporous material can also be used at the bottom
of the cell, through which the gas bubbles are produced. Zhou
et al. (1994, 1995, 1997) have also proposed a method of gener-
ating tiny bubbles by hydrodynamic cavitation, which is the pro-
cess of creation and growth of gas bubbles in a liquid due to the
rupture of a liquidliquid or a liquidsolid interface under the
inuence of external forces. The bubbles generated on a particle
surface by cavitation naturally attach to the particle, eliminating
the collision and attachment process, which is often the rate-
determining step for otation. Cavitation also improves the ota-
tion efciency of coarse particles by reducing the detachment
probability during the rise of particlebubble aggregates in the
liquid.
Physiochemical methods include dissolved air otation and
electrootation. These methods are reviewed below.
2.1.1. Dissolved gas otation
One technique to generate small bubbles and to improve the
oatability of ne particles is dissolved gas otation (Matis and
Gallios, 1986; Matis et al., 1993). The most common dissolved
gas is air so the name dissolved air otation is more commonly
used. Dissolved air otation is based on Henrys law, where the
solubility of air in an aqueous solution is proportional to the par-
tial pressure of air at constant temperature. By subjecting the
solution to an over-pressure, air molecules dissolve in the solu-
tion. The solution becomes supersaturated when the pressure is
released and gas molecules precipitate as small bubbles. Another
way to precipitate bubbles is to decrease the pressure by applying
a vacuum, a technique called vacuum otation. The gas bubbles
nucleate preferentially on the hydrophobic surfaces of minerals,
thus eliminating the bubbleparticle collision step in conven-
tional otation. The air trapped in pores and crevices could also
act as a growth centres for bubbles. As pores and crevices may
be common to the minerals to be oated as well as the gangue,
the growth of the air pockets in pores and crevices is detrimental
in general.
The efciency of dissolved air otation can also be accounted
for by a two-stage otation process, which is a combination of
dissolved air otation and conventional otation. Dziensiewicz
and Pryor (1950) and Klassen and Mokrousov (1963) proposed that
the adsorption of nuclear air on mineral particles during dis-
solved air otation aided the attachment of these particles to con-
ventional sized bubbles. The recovery is mainly improved because
the attachment of a large bubble to a mineral particle by coales-
cence with a frosting of tiny bubbles on the particle is more favour-
able thermodynamically compared with direct attachment
(Mishchuk et al., 2002, 2006). Thus, bubblebubble coalescence is
more favoured than bubbleparticle attachments, since the van
der Waals interaction is attractive in the former case.
Dai et al. (1998) demonstrated that different gases exert con-
trasting behaviour in otation. Carbon dioxide and air play a signif-
icant role in enhancing particlebubble capture. The dissolution of
other gases such as helium, argon, hydrogen, nitrogen and oxygen
decreased the concentration of carbon dioxide in solution and
inhibited the particlebubble capture mechanism. A mechanism
by which CO
2
is generated in the intervening thin aqueous lm be-
tween a hydrophobic particle and a bubble, leading to an increased
capture efciency, was proposed.
A problem with dissolved air otation, at least where mineral
processing is concerned, is that the volume of gas produced is far
lower than with mechanical systems. The air ow rate is typically
10 m
3
/m
3
pulp in a mechanical cell, whereas in dissolved air ota-
tion it is about 0.007 m
3
air/m
3
water (Matis, 1995). These values
equal approximately 6450 min
1
and 60 min
1
in the supercial
surface area rate of bubbles, respectively. Another problem is that
microbubbles cause high water recoveries (Schwarz and Grano,
2002). It is observed that entrainment of gangue minerals is pro-
portional to the recovery of water (Trahar and Warren, 1976; Liu
and Wannas, 2004). Therefore, dissolved air otation is mainly ap-
plied in wastewater treatment and industrial efuent treatments,
where selectivity is not needed (Matis, 1995).
2.1.2. Electrootation
The application of electrootation for the recovery of ne min-
erals has been under investigation since 1946 when it was rst ap-
plied to mineral beneciation in the USSR (Matis and Backhurst,
1984). Its application to ne mineral recovery has become a topic
of recent research, due to a combination of uneconomic and inef-
cient conventional ne mineral otation techniques accompanied
by an increased demand for minerals (Matis and Backhurst, 1984).
Electrootation is based on the use of hydrogen and oxygen
bubbles formed by the electrolysis of water. Oxygen is formed at
the anode, Eq. (1), and hydrogen is formed at the cathode, Eq.
(2). These two gases may be used either separately or together,
or in combination with air bubbles in electrootation
H
2
O 2H

1
2
O
2
(g) 2e

(1)
2H
2
O 2e

2OH

H
2
(g) (2)
In most cases electrolytic oxygen and hydrogen do not have
any specic effect on the surface of minerals and their otation
rate depends mainly on the bubble size (Glembotskii et al.,
1975). For sulphide minerals, Glembotskii et al. (1975) showed
that they undergo physico-chemical changes in an electrootation
cell, inuencing their oatability. They performed otation exper-
iments with pyrite in the presence of potassium xanthogenate and
showed that bubbles of electrolytic oxygen caused a 98% recovery
of the mineral, whereas hydrogen achieved practically no otation
at all. They detected an increase in the concentration of Fe
2+
and
SO
4
2
ions in the pulp and elementary sulphur on the mineral
particle surfaces. They also suggested that electrolytic oxygen
makes the surfaces of the pyrite and other sulphide (chalcopyrite,
sphalerite) particles so strongly hydrophobic that there is no need
T. Miettinen et al. / Minerals Engineering 23 (2010) 420437 421
to use a otation reagent. They also pointed out that intensive
oxidation of sulphides in conditions of electrootation does not al-
ways lead to an increased otation activity of minerals. The posi-
tive effect of oxygen on the otation of ne particles of
chalcopyrite has also been reported elsewhere (Raju and Kha-
ngaonkar, 1982, 1984).
Electrootation has mostly been applied to water and wastewa-
ter treatment. More specically, it has been used for the reduction
of suspended solids, phosphates and dissolved metals as well as for
sludge thickening (Matis and Zouboulis, 1995; Romanov, 1998).
2.1.3. Turbulent microotation
The bubbles in conventional otation machines are typically
0.52 mm in diameter (Gorain et al., 1997). Rulyov (2001) demon-
strated the effectiveness of small bubbles in a microotation pilot
plant. The theory of turbulent microotation is based on pre-aggre-
gation of particles to reach an aggregate size of greater than 7 lm,
utilisation of microbubbles with initial dimensions smaller than
40 lm and maintaining the conditions for microbubble coales-
cence efciency so that the latter is smaller than 0.03. Microota-
tion is performed in a long, narrow channel, where the treated
mixture ows after saturation with gas microbubbles. The ow
rate of the mixture is adjusted in such a way as to maintain a tur-
bulent stream ow and, secondly, to ensure that aggregation of
suspended particles, heterocoagulation of particles and bubbles
and aggregation of bubbles with particles attached to them take
place. The otation times that Rulyov obtained in his pilot plant
experiments were around 10-fold lower compared with dissolved
air otation. An outline of Rulyovs microotation pilot plant is
shown in Fig. 1.
2.2. Enhancing particleparticle interaction
For ne particles, it is experimentally and theoretically clear
that the otation rate increases with increasing particle size (Suth-
erland, 1948; Trahar and Warren, 1976; Crawford and Ralston,
1988; Ralston and Dukhin, 1999; Duan et al., 2003; Pyke et al.,
2003). Thus, many techniques have been developed which try to
increase particle size and mass and decrease surface energy. All
of these techniques have the same feature; ne particles are in-
duced to form ocs or aggregates. Depending on the mechanisms
of aggregate formation, these techniques can be divided into three
classes: selective occulation, coagulation and hydrophobic
aggregation.
In selective occulation, the ocs are formed due to the bridging
ability of long-chain polymer molecules or ions. These polymers
rst adsorb onto mineral surfaces by electrostatic forces, specic
chemical interaction and/or hydrogen bonds followed by bridging
other mineral particles and forming loose ocs (Gregory, 1989).
It has often been claimed that selective occulation is a promising
technique for ne mineral particles. Several experimental results
have showed improved otation of ne particles by selective oc-
culation (Attia, 1977; Song et al., 2000). However, as indicated by
Rubio et al. (2003), selective occulation is yet to nd general
application, for entrapment of gangue is a major issue.
Coagulation of ne particles can be achieved by the addition of
electrolyte which decreases the electrostatic repulsion between
charged particles. Coagulation is a much-used technique in water
purication where no selectivity is required. However, in the min-
erals processing industry, selectivity is essential. Electrolyte addi-
tion often causes heterocoagulation, however it is very hard to
obtain selective coagulation only by electrolyte addition. There-
fore, this method is rarely used in the minerals processing industry
for particle size enlargement.
With hydrophobic aggregation, the ne particles are selectively
hydrophobised, similar to conventional froth otation. In order for
the particles to be held together by hydrophobic forces they need
to be in very close proximity, achieved by intense agitation (Koh
and Warren, 1980; Warren, 1984, 1992). Non-polar oil is often
added to the solution to increase the strength of the aggregates.
Hydrophobic aggregation can be further divided into shear occu-
lation, emulsion otation, carrier otation, oil extended otation,
spherical agglomeration and two liquid extraction (Lai and Fuer-
stenau, 1968; Hoover and Malhotra, 1976; Fuerstenau, 1980;
Subrahmanyam and Forssberg, 1990; Warren, 1992; Rao, 1997).
Floc otation has been found to produce much higher chalcopy-
rite recovery, for example, than conventional otation (Peng et al.,
2005). However, like microbubble otation, oc otation also pro-
duces high entrainment/entrapment of gangue minerals (Attia,
1982). Liu and Wannas (2004) suggest that to reduce and eventu-
ally to eliminate mechanical entrainment, the particle size of the
hydrophilic particles might also be enlarged, by polymer depres-
sants. They realised that otation depressants make minerals
hydrophilic so that they do not oat, but that a hydrophilic mineral
particle can be collected into the froth by mechanical entrainment
if it is sufciently small. Therefore, they proposed that selective
depressants should not only render mineral particles hydrophilic,
but also occulate the hydrophilic particles, thus reducing their
entrainment as well. Peng et al. (2005) found that by using suitable
polymeric depressants that also possessed occulating power, it
was possible to reduce the mechanical entrainment of the ne
hydrophilic minerals without affecting the otation of the hydro-
phobic value.
Fig. 1. The pilot plant (scheme) for turbulent microotation of alumina. (1) Elec-
trolyser (microbubble generator); (2) tubular static mixer; (3) samplers; (4)
microphotography cell; (5) foam separator; (6) otosludge collector; (7) disk foam
breaker; (8) otosludge outlet (Rulyov, 2001). Aethonium (C
30
H
62
Cl
2
O
4
) was used
to ensure aggregation of alumina, attachment of alumina to bubbles and aggrega-
tion of bubbles and potassium iron hexacyanoferrate was used for visualisation.
Reproduced with permission from Elsevier, Rulyov, N.N., Turbulent microotation:
theory and experiment, Colloids and Surfaces A: Physicochemical and Engineering
Aspects 192(13) (2001) 7391.
422 T. Miettinen et al. / Minerals Engineering 23 (2010) 420437
3. Flotation kinetics
Theory and experiment indicate that the otation process is
rst-order with respect to the number of particles, N (Sutherland,
1948; Jameson et al., 1977). Thus, the rate equation for the removal
of particles in a batch process is of the form
dN
dt
= kN (3)
where k is the otation rate constant and t is the otation time. If
the initial number of particles is N
0
at t = 0, Eq. (3) can be integrated
to yield
N = N
0
e
kt
(4)
The recovery of the particles, R, is dened by
R =
N
0
N
N
0
(5)
In terms of the recovery, Eq. (4) becomes
R = R
max
(1 e
kt
): (6)
where R
max
is the recovery after innite time.
In a simple batch otation case where mixing is not involved
(Jameson et al., 1977; Yoon and Mao, 1996), the otation rate con-
stant is
k =
3J
g
2d
b
E
Col
(7)
where d
b
is the bubble diameter, E
Col
is the bubbleparticle collec-
tion efciency and J
g
is the supercial gas velocity, dened as volu-
metric gas ow rate divided by the cross-sectional area of the
otation column. The value of k is normally determined experimen-
tally from a plot of ln(1 R) versus time. In the case of monodi-
spersed particles with the same surface hydrophobicity, a linear
plot is obtained and k is determined from the slope. In the case of
polydisperse particle size systems, a two-component t is used
and fast, k
f
, and slow, k
s
, rate constants are obtained (Ralston, 1992).
The capture of a particle and a bubble is generally divided into
three separate processes (Derjaguin and Dukhin, 1961). Firstly, the
particle is subjected to hydrodynamic forces far away from the
bubble surface. Hydrodynamic drag forces act to sweep the particle
around the bubble surface. Viscous forces retard this relative mo-
tion whilst particle inertial and gravity forces drive the particle to-
ward the bubble surface. Secondly, the surface forces between the
bubble and particle have to favour thin lm rupture and the forma-
tion of a three-phase contact line in order for the particle to attach
to the bubble. Thirdly, the formed bubbleparticle aggregate has to
be stable, i.e., the attachment forces between the bubble and par-
ticle have to be larger than the detachment forces. Particles which
satisfy these conditions can be separated selectively from gangue
particles which fail one or more of these conditions. The bubble
particle collection or capture efciency, E
Col
, can be dened as a
product of bubbleparticle collision, E
C
, attachment, E
A
, and stabil-
ity, E
S
, efciencies, since these processes, all of which are probabil-
ities, are independent of each other. This dissection was formally
proposed by Derjaguin and Dukhin (1961)
E
Col
= E
C
E
A
E
S
(8)
4. Bubbleparticle collision
Before bubbleparticle attachment can occur, a particle has to
collide with a bubble, reaching a separation distance at which sur-
face forces start to operate (Schulze, 1984). The various bubble
particle collision mechanisms and models are described and the
factors inuencing bubbleparticle collision are summarised.
4.1. Collision mechanisms
The determination of bubbleparticle collision involves the
evaluation of forces that cause a particle to deviate in its trajectory
from uid streamlines near the bubble surface and collide with a
bubble. The forces that affect the motion of particles include grav-
itational forces, inertial forces and hydrodynamic drag forces. Bub-
bleparticle collision mechanisms by diffusion (Reay and Ratcliff,
1973; Yang et al., 1995; Nguyen et al., 2006) and shear (Abraham-
son, 1975) have also been dened. Collision by shear-induced
mechanisms is usually not considered in otation since they are
only signicant for the collision of spheres of similar sizes. The ef-
fect of uid distortion by the spheres on the bubbleparticle colli-
sion efciency is negligibly small (Nguyen and Schulze, 2004).
Four bubbleparticle collision mechanisms, involving inertia,
gravity, interception and Brownian diffusion are shown in Fig. 2.
The inertial collision mechanism is most likely for coarse and dense
particles which are unable to follow uid streamlines and tend to
move along a straight path. If the density of particles is greater
than that of the surrounding uid, particles have a certain settling
velocity and therefore their trajectory deviates from uid stream-
lines. This deviation may cause particles to collide with the bubble
surface. The collision of particles with the bubble surface by inter-
ception is due to a ow which carries particles along the uid
streamlines. The particles come into contact with the bubble sur-
face because of their nite size. Bubbleparticle collision by
Brownian diffusion is signicant for submicron particles which
move randomly in the uid.
Bubbleparticle collision may occur by the individual mecha-
nisms described above or it could be a result of two or more
of these mechanisms. According to Derjaguin et al. (1984), the
mechanism of transfer of small particles to the bubble surface is
Fig. 2. Schematic representation of (a) inertia, (b) gravity, (c) interception and (d)
Brownian collision mechanisms. The thick lines represent particle trajectories
whilst the thin lines represent the uid streamlines.
T. Miettinen et al. / Minerals Engineering 23 (2010) 420437 423
mainly governed by two parameters: the inertial forces and the
long-range hydrodynamic interaction (LRHI), i.e., hydrodynamic
drag forces. As described above, the inertial forces dominate the
LRHI in the case of large particles, and the smaller the particle is
and the closer the density of the particle is to the density of sur-
rounding uid the more the LRHI changes the trajectory of the par-
ticle to coincide with the uid streamlines. The ratio of the
stopping distance of the particle in the presence of the LRHI to a
characteristic dimension of the obstacle is called the Stokes
number
St =
q
p
u
b
d
2
p
9d
b
g
f
(9)
where u
b
is the velocity of the bubble, g
f
is the dynamic viscosity of
the uid and q
p
and d
p
are the density and the diameter of the par-
ticle, respectively. This dimensionless number can be used to char-
acterise the shape of the particle trajectory in the uid ow and the
interactions between the particle and bubble surfaces (Ralston et
al., 2002). For conditions where:
v St 1: inertial forces have practically no effect on the motion of
the particles, which can be considered as inertia-free.
v St 6 0.1: negative inertial forces can impede particle deposi-
tion on a bubble (Dai et al., 1998).
v 0.1 < St < 1: an inelastic inertial impact of particles on a bubble
surface is characteristic of this regime and a major portion of
the kinetic energy of the particles is lost both during the
approach to the bubble and at the impact itself, when a liquid
layer is formed between the surfaces of a particle and a bubble.
v St > 3: the trajectory of a particle deviates very slightly from a
straight line and the energy of the particle, as it approaches
the bubble and on collision, changes so little that the impact
can be considered as being quasi-elastic, i.e., the particle
bounces away from the bubble surface at almost the same speed
as it approaches the bubble surface.
Another important dimensionless number which is generally
used to assess the ow conditions of uid in proximity to various
objects, is the Reynolds number, Re
b
. The Reynolds number is the
ratio of the inertial forces to the viscous forces of the uid
Re
b
=
u
b
q
f
d
b
g
f
: (10)
where q
f
is the density of the uid. The uid streamlines around
bubbles can be calculated using the analytical solutions of the con-
tinuity equations at two extreme ow conditions, i.e., Stokes and
potential ow. The Stokes ow conditions apply when the bubble
Reynolds number is very much less than unity and potential ow
conditions apply at 80 < Re
b
< 500 (Schulze, 1993).
4.2. Fluid streamlines
The process of collecting particles by bubbles generally takes
place in a complex and highly turbulent environment inside a o-
tation cell. The particle and bubble motions in turbulent conditions
are complicated, for turbulence causes bubbles and particles to
move relatively to each other as well as to the uid surrounding
them. Secondly, the spatial variations, which are an essential fea-
ture of turbulent motions, can cause small particles to diffuse from
one point to another by eddy motion (Nguyen and Schulze, 2004).
The turbulent collision problem has been analysed by several
authors (Levich, 1962; Abrahamson, 1975; Jordan and Spears,
1990; Rulyov, 2001; Pyke et al., 2003). However, the liquid stream-
lines around the bubble may be described under more quiescent
conditions (e.g. potential ow) as the distance separating the par-
ticle and the bubble decreases. The transport of particles towards
bubble surfaces at short surface to surface distances, where drain-
age processes are dominant, may be uninuenced by bulk turbu-
lence effects (Dai et al., 1998, 2000). In the case of column
otation, where external mixing is not used, rather quiescent con-
ditions are encountered (Yoon and Luttrell, 1989).
The uid streamlines around spherical objects under various
ow conditions can be described by stream functions. In the case
of the Stokes ow condition, i.e., Reynolds numbers much less than
unity, the stream function can be determined analytically by solv-
ing the NavierStokes equation. The derivation of these equations
can be found in standard textbooks on uid dynamics (Acheson,
1990; Johnson, 1998; Batchelor, 2000). The Stokes ow stream
function
W
St
= u
b
R
2
b
sin
2
h
1
2
x
2

3
4
x
1
4x
_ _
(11)
describes the uid streamlines in polar coordinates, where h is the
angular component, r is the radial component, R
p
and R
b
are the ra-
dii of the particle and bubble, respectively, and x = r/R
b
(Fig. 3).
In the inviscid ow case, where the bubble Reynolds number is
very high, the stream function can be approximated by the poten-
tial ow solution of the NavierStokes equation,
W
pot
= u
b
R
2
b
sin
2
h
1
2
x
2

1
2x
_ _
(12)
The bubble size range in otation cells usually falls between the
Stokes and potential ow conditions. Yoon and Luttrell (1989)
Fig. 3. Particle of radius R
p
approaching a bubble of radius R
b
along uid
streamlines. See text for explanation.
424 T. Miettinen et al. / Minerals Engineering 23 (2010) 420437
determined a stream function for the intermediate ow conditions,
Eq. (13), by combining Eqs. (11) and (12)
W
YL
= u
b
R
2
b
sin
2
h
1
2
x
2

3
4
C
1
x
1
2x

3C
1
4x
_ _
(13)
where C
1
is a dimensionless parameter which varies between 0 and
1. If the value of C
1
is equal to 0, the potential ow stream function
applies, whereas if C
1
is equal to 1, the Stokes ow stream function
applies. The values of C
1
were determined by tting streamline data
from the literature. Yoon and Luttrell found that C
1
values could be
normalized with respect to the bubble Reynolds numbers by using
Eq. (14) for bubble Reynolds numbers from 1 to 100
C
1
= 1
4Re
0:72
b
45
1
1
x
_ _
(14)
The streamlines are the trajectories that a small, neutrally buoy-
ant particle will follow during an encounter with a bubble. Such a
particle will collide with the bubble only if it is on a streamline that
has its closest approach to the bubble equal to, i.e., a grazing trajec-
tory, or less than the radius of the particle. The former streamline
passes through the point (r = R
p
+ R
b
, h = /2; see Fig. 3), which en-
ables one to determine the critical stream function for a bubble
particle collision. When the particle is at an innite distance above
the bubble, then h = 0 for the critical stream function and it is pos-
sible to obtain the maximum distance, at which the centre of the
particle can lie from the centre of the line of motion of the bubble
in order to just collide with it, R
cr
. All particles of similar size lying
within this distance from the line of motion of the bubble will col-
lide with it and therefore the bubbleparticle collision efciency,
E
C
, is determined by the ratio of the cross-sectional area pR
2
cr
and
the projected area of the bubble and the particle
E
c
=
pR
2
cr
p(R
p
R
b
)
2
: (15)
4.3. Collision models
A detailed review of bubbleparticle collision models has been
given by Dai et al. (2000). The most relevant collision models for
ne particle otation are reviewed here, including recent models.
The rst ve models involve purely bubbleparticle collision,
whereas the last three models are mixtures of bubbleparticle col-
lision and collection, for bubbleparticle attachment is incorpo-
rated into these models. Some authors also regard bubble
particle collision by Brownian diffusion as bubbleparticle
collection.
4.3.1. Stokes ow model
Gaudinet al. (1942) andGaudin(1957) assumeda Stokes owre-
gime around the bubble surface and ignored the inertial force of the
particles. Using the Stokes ow stream function, Eqs. (11) and (15),
Gaudin obtained the Stokes ow bubbleparticle collision model
E
C-St
=
3
2
d
p
d
b
_ _
2
(16)
The Stokes ow bubbleparticle collision model applies when
the bubble Reynolds number is much less than unity (Schulze,
1993).
4.3.2. Potential ow model
Sutherland (1948) developed one of the earliest collision mod-
els between a single particle and a single bubble, applying the prin-
ciples of interception mentioned above. He used the potential ow
conditions to calculate the streamlines around the bubble and the
bubbleparticle collision efciency, E
C-pot
. He was aware of Brown-
ian motion and its effects on the bubbleparticle collision (Reay
and Ratcliff, 1973; Jameson et al., 1977; Yang et al., 1995; Nguyen
et al., 2006) and he set a condition that the size of the particle (or
bubble) must be larger than 0.1 lm
E
C-pot
=
3d
p
d
b
(17)
The potential ow bubbleparticle collision model applies
when the bubble Reynolds number is between 80 and 500 (Schu-
lze, 1993).
4.3.3. Yoon and Luttrell ow model
Yoon and Luttrell (1989) derived a bubbleparticle collision
model, applying the interception mechanism and an empirical
stream function Eq. (13), which is valid for intermediate ow con-
ditions, i.e., between the potential and the Stokes ow conditions
E
CYL
=
3
2

4Re
0:72
b
15
_ _
d
p
d
b
_ _
2
(18)
The Yoon and Luttrell bubbleparticle collision model applies
when the bubble Reynolds number is between 1 and 100 (Yoon
and Luttrell, 1989).
4.3.4. Generalised Sutherland equation
Sutherlands equation is a reasonable description when poten-
tial ow conditions exist and the particle size is close to its lower
limit of 0.1 lm, but when the particle size is larger, inertial and
gravity effects have to be taken into account.
The generalised Sutherland equation (GSE) (Dai et al., 1998),
E
C-GSE
, couples the interception mechanism of the Sutherland mod-
el to the positive effect of the hydrodynamic pressing force and the
negative effect of centrifugal force. The GSE can be applied to the
potential ow conditions for mobile bubble surfaces and for Stokes
numbers less than 0.1. This Stokes number regime allows the par-
ticle to deviate from the streamline but not enough to cause any
difference in the acceleration of the particle from that of the uid
E
c-GSE
= E
c-pot
sin2ht exp(I
1
I
2
): (19)
In Eq. (20), h
t
is the angle of tangency, I
1
is the positive effect of
the hydrodynamic pressing force and I
2
is the negative effect of
centrifugal force, dened in Eqs. (21) and (22). The total inertial
force is the sum of the hydrodynamic pressing force and the cen-
trifugal force
I
1
= 3K
3
ln
3
E
C-pot
1:8
_ _
cos h
t
(20)
I
2
=
4cos h
t
2
3

cos
3
ht
3
cos h
t
_ _
sin
4
h
t
(21)
In the distant part of the particle trajectory, Fig. 4, the particle
moves toward the bubble surface under the action of the normal
component of the liquid ow. As the particle approaches close to
the bubble surface, the drainage of the liquid lm between the
bubble and particle occurs, a hydrodynamic resistance develops,
and the particle movement slows down compared with the normal
component of the liquid velocity. The particle is subjected to a net
positive force, due to this velocity difference, termed the hydrody-
namic pressing force. This is insufcient to ensure contact between
the particle and the bubble because the hydrodynamic resistance
increases indenitely as the liquid layer becomes thinner. The
hydrodynamic pressing force is proportional to the difference be-
tween the radial components of the liquid and particle velocities.
The tangential velocity is the origin of the centrifugal forces which
act to prevent particle deposition near the equator.
T. Miettinen et al. / Minerals Engineering 23 (2010) 420437 425
The effect of inertial forces on the particle trajectory can be seen
in Figs. 4 and 5. In the distant part of the particle trajectory the to-
tal inertial force is always positive and thus improves the collision
efciency. In the near part of the particle trajectory the centrifugal
force causes the particle to deviate away from the bubble surface
and thus decreasing the collision efciency. In Eqs. (20) and (23)
and in Figs. 4 and 5, h
t
is the angle of tangency, i.e., the angle where
the total normal force is zero. At h < h
t
the pressing force is domi-
nant and the particle deposition is enhanced. At h > h
t
the negative
centrifugal force impedes the collision. This means that the particle
moving along the grazing trajectory touches the bubble surface
just at the angle h
t
. The angle of tangency is dened as
h
t
=
360
2p
arcsin(2b((1 b
2
)
1=2
b))
1=2
: (22)
where
b =
4E
C-pot
9K
3
(23)
and
K
1
= St 1
q
f
2q
p
_ _
(24)
K
2
= 3St
q
f
2q
p
(25)
K
3
= K
1
K
2
(26)
where b, K
1
, K
2
and K
3
are dimensionless numbers.
4.3.5. Rulyov model
Rulyov (1989, 2001) introduced interception collision mecha-
nism for bubble diameters less than 600 lm in his turbulent
microotation theory. He assumed that these sizes of bubbles are
normally retarded by the adsorbed surfactant layer and dened
the bubbleparticle collision efciency as
E
C-R
= C(Re
b
)
d
p
d
b
_ _
2
(27)
where Re
b
is the bubble Reynolds number, d
p
and d
b
are the particle
and bubble diameters, respectively. This equation can be applied
over a range of bubble Reynolds numbers between 1 and 40, where
C(Re
b
) 1.5 + 3650l
2
b
. Thus this model is an intermediate ow mod-
el, where C(Re
b
) term takes a change in ow conditions on the bub-
ble surface into account. By examining Eq. (28), it can be seen that if
the bubble velocity is higher than rst-order with respect to the
bubble diameter (usually second-order), E
C-R
increases with increas-
ing bubble size. This behaviour is opposite to what can be seen in
the other collision models reviewed here.
4.3.6. Reay and Ratcliff model
Reay and Ratcliff (1973) expected the bubbleparticle collection
efciency to have a minimum, because colloidal particles reach the
bubble mainly by Brownian diffusion while particles larger than
about 3 lm have a better chance to be intercepted by the bubble
as the particle size increases. They derived a collection model for
particles less than 0.2 lm in diameter and a collision model for
particles between 3 and 20 lm in diameter, assuming Stokes ow
conditions around the bubble.
For the collection model in the diffusion regime Reay and Rat-
cliff dened C
B
as the concentration of particles in the bulk liquid
and C
S
as the concentration of unadsorbed particles in the layer
adjacent to a bubble surface. If the particles are adsorbing strongly,
C
S
is equal to 0 and if the particles are adsorbing slowly C
S
is equal
to C
B
. They expressed the particle diffusivity D
p
by the StokesEin-
Fig. 4. Schematic representation of the point of inection which divides a
streamline into the distant part and the near part. Under the action of inertial
forces, the particle trajectory deviates from the uid streamline toward the bubble
surface in the distant part, but away from the bubble surface in the near part (Dai
et al., 1998). See text for explanation. Reproduced with permission from Elsevier,
Journal of Colloid and Interface Science 217(1) (1999) 7076.
Fig. 5. Schematic representation of the angular dependence of the pressing force and the centrifugal force (a) and the total force (b) at the bubble surface (Dai et al., 1998).
Reproduced with permission from Elsevier, Journal of Colloid and Interface Science 217(1) (1999) 7076.
426 T. Miettinen et al. / Minerals Engineering 23 (2010) 420437
stein relation Eq. (29), where k
B
is the Bolzman constant, g
f
is the
dynamic viscosity of the uid, R
p
is the particle radius and T is
the absolute temperature, in order to derive the particle mass
transfer coefcient k
p
for Ficks law
D
p
=
k
B
T
6pg
f
R
p
(28)
In Stokes ow conditions, the concentration boundary layer
theory yields
Sh ~ Pe
1=3
(29)
where Sh is the Sherwood number, Eq. (31), and Pe is the Peclet
number, Eq. (32)
Sh =
2R
b
k
p
D
p
(30)
Pe =
2R
b
u
b
D
p
(31)
The net ow of particles to the bubble surface in unit time, by
Ficks law, is given in Eq. (33) and the number of particles in a vol-
ume, which the bubble sweeps in unit time, is given in Eq. (34)
N
//
= 4pRb2kp(C
B
C
S
) (32)
N
/
= pR
2
b
u
b
C
B
: (33)
Reay and Ratcliff derived the collection efciency, E
Col-RR
, as a
ratio of Eqs. (33) and (34). They expressed the bubble rising veloc-
ity as the Stokes terminal velocity and obtained Eq. (35) at 25 C
E
Col-RR
= 1:17 10
11
C
B
C
S
R
2
b
R
2=3
p
C
B
(34)
This concentration boundary layer theory is valid when the
average thickness of the concentration boundary layer is much
greater than the size of the particle. Reay and Ratcliff concluded
that Eq. (35) can only be applied when the particle diameter is less
than 0.2 lm, and when the average thickness of the concentration
boundary layer is about 1.6 lm. Eq. (35) shows that the bubble
particle collection efciency increases with decreasing particle
and bubble size.
In their collision model for particles between 3 and 20 lm, they
proposed that gravity is the only factor causing the particles tra-
jectory to deviate signicantly from the uid streamlines. They
developed a model, where the bubbleparticle collision efciency
was proportional to the dimensions of the bubble and particle
and the densities of the liquid and particle. For particle to liquid
density ratios of 1.0 and 2.5, Eqs. (36) and (37) were obtained,
respectively
E
CRR
= 1:25
R
p
R
b
_ _
1:9
(35)
E
CRR
= 3:6
R
p
R
b
_ _
2:05
(36)
Over this density range E
C-RR
is roughly proportional to (R
p
/R
b
)
2
,
which is the same as in the case of the Stokes ow model. Reay and
Ratcliff (1973, 1975) also measured the otation rates of glass
beads and latex particles using very small bubbles. Their collision
model only worked reasonably well for electrically uncharged par-
ticles, which indicated that the interaction forces between surfaces
have to be taken into account. The same problem exists in their dif-
fusion model, where the effect of surface forces is even more pro-
nounced, because the particles are almost weightless.
4.3.7. Collins model
Collins (1975) obtained the bubbleparticle collision efciency
by determining the concentration of particles near to the bubble
and hence the ux onto the bubble. Integration of the ux over
the bubble yielded the bubbleparticle collision efciency. Collins
used a hydrodynamic correction in his model following the ap-
proach of Spielman and Goren (1970) and Prieve and Ruckenstein
(1974). The relationship between the bubbleparticle collision ef-
ciency and the bubble and particle size was described as
E
C-C

1
R
2
b
R
2=3
p
(37)
It can be seen that the relationship is the same as in the Reay
and Ratcliff model in Eq. (35). The predicted bubbleparticle colli-
sion efciency shows a minimum at about 0.6 lm in agreement
with the results of Reay and Ratcliff (1973) but the magnitude is
higher, because of the hydrodynamic correction (Jameson et al.,
1977).
4.3.8. Yang et al. model
Yang et al. (1995) also considered the Brownian motion of
small particles toward the bubble surface, as well as the capture
by interception due to the motion of small particles relative to
the rising bubble, in a similar fashion to Reay and Ratcliff (1973).
In their theoretical study they determined the bubbleparticle col-
lection efciencies by means of the Stokes and potential ow
stream functions. Furthermore they included Marangoni effects
and hydrodynamic corrections. The Marangoni effect happens
when surfactants or small particles are adsorbed on the bubble
surface and are then swept along the interface towards the rear
of the bubble. The resulting inhomogenity in surfactant concentra-
tion acts to create interfacial tension gradients along the bubble
surface which drives the system back toward a uniform surfactant
concentration (Davis and Acrivos, 1966). The Marangoni number,
Ma, is dened in Eq. (39)
Ma =
E
0
aR
b
g
f
(38)
where a is the adsorption parameter of particles at the equilibrium
state and E
0
is the Gibbs elasticity for the surface active particles.
The Marangoni number represents the ratio of interfacial tension
gradient forces to viscous forces (Edwards et al., 1991). Therefore,
the bubble surface is fully retarded in the case of Ma ?and mo-
bile with nite Marangoni numbers.
Yang et al. (1995) based their diffusive particle transport model
on the Peclet number Eq. (32) and the StokesEinstein Eq. (29). In
the Stokes ow conditions, the collection efciencies were de-
scribed for a mobile bubble surface by Eq. (40) and for a retarded
bubble surface by Eq. (41)
E
Col-Y1
=
1:842

1
2
3
Ma
_ _
Pe
_
R
p
R
b
Pe
(39)
E
Col-Y2
= 2:498Pe
2=3

R
p
R
b
Pe
(40)
where R
p
/(R
b
/Pe) is the hydrodynamic correction term. In practice,
the hydrodynamic correction term can be neglected in Eqs. (40)
and (41), because the particle diffusivities are about thousand times
smaller than molecular diffusivities, which makes the Peclet num-
ber large and the hydrodynamic correction term very small com-
pared with the diffusion term. In the potential ow regime, the
diffusive collection efciency Eq. (42) was only derived for the mo-
bile surface, because the potential ow solution does not satisfy the
fully retarded bubble surface condition
T. Miettinen et al. / Minerals Engineering 23 (2010) 420437 427
E
Col-Y3
= 3:192Pe
1=2

R
p
R
b
Pe
(41)
In the interception domain, they derived the bubbleparticle
collection efciency classically by assuming that the particle fol-
lows uid streamlines around the bubble. As a result, they ob-
tained Eqs. (43) and (44) under Stokes ow conditions, where
the former is for a mobile bubble surface and the latter is for the
fully retarded bubble surface
E
Col-Y4
=
1
1
2
3
Ma
R
p
R
b
_ _
1 Ma ( )
R
p
R
b
_ _
2

R
p
R
b
_ _
3
_ _
: (42)
E
Col-Y5
=
3
2
R
p
R
b
_ _
2

R
p
R
b
_ _
3
: (43)
In the case of the potential ow conditions, Eq. (45) applies
E
Col-Y6
= 3
R
p
R
b
_ _

R
p
R
b
_ _
3
_ _
: (44)
Eqs. (43) and (45) have been determined without considering
any hydrodynamic interactions between the particle and the bub-
ble. In a subsequent analysis, Yang et al. (1995) determined the
hydrodynamic correction to these collection efciencies. Their re-
sults indicated that the contribution to the collection efciency
from the hydrodynamic interactions is negative and that the inter-
actions become stronger as the particle size increases. The hydro-
dynamic interactions are also stronger in the case of potential
ow than in the case of Stokes ow.
Yang et al. (1995) determined the overall bubbleparticle col-
lection efciency as a sum of the diffusion and interception contri-
butions. They showed that the collection efciency decreases in the
diffusion region and increases in the interception region with
increasing particle size. In turn, a mobile bubble surface and
decreasing bubble size increased the collection efciency. Yang
et al. (1995) also calculated the critical particle size where the par-
ticles transport mechanism changes from diffusion to interception.
A minimum of the collection efciency was roughly at a particle
diameter between 0.1 and 0.2 lm in the case of the mobile bubble
surface and 0.45 lm in the case a fully retarded bubble surface. The
critical particle size decreased with increasing bubble size.
4.3.9. Nguyen et al. model
Nguyen et al. (2006) presented a theoretical model for bubble
particle collection by Brownian diffusion and convection. The mod-
el incorporated fundamental theories of mass transfer by diffusion,
convection and migration, microhydrodynamics and colloidal
forces. The emphasis on microhydrodynamics was focused on a
mobile bubble surface. The colloidal forces included the van der
Waals, electrostatic double layer and hydrophobic forces. The
equation was solved for a particle concentration distribution,
employing a numerical computation method. The bubbleparticle
collection efciency was numerically validated as a function of
particle size, zeta potential and hydrophobic force constants and
decay lengths against the experimental otation results obtained
with 40160 nm quartz particles. The theoretical and experimental
results showed that the bubbleparticle collection efciency has a
minimum at a particle diameter in the order of 100 nm. For larger
particles, the interception and collision mechanisms predomi-
nated, while the diffusion and colloidal forces controlled the collec-
tion of particles with a size smaller than the transition size.
4.4. Factors inuencing bubbleparticle collision
Various parameters inuence bubbleparticle collision ef-
ciency. These parameters include particle size and density, bubble
size, bubble rising velocity and bubble surface mobility.
Increasing bubbleparticle collision efciency is predicted by all
the collision models reviewed above with increasing particle size
and decreasing bubble size (the Rulyov model excluded). This is
also in agreement with bubbleparticle collision experiments (Anf-
runs and Kitchener, 1977; Yoon and Luttrell, 1989; Hewitt et al.,
1995; Dai et al., 1998). This is valid for particles where Brownian
motion is unimportant. For submicron particles, where Brownian
motion is the major collision mechanism, the bubbleparticle col-
lision efciency has been found to increase with decreasing parti-
cle and bubble size (Reay and Ratcliff, 1973; Yang et al., 1995;
Nguyen et al., 2006). This has also been observed in experiments
(Nguyen et al., 2006).
The inuence of the density of particles and bubble rising veloc-
ity can be seen in the Stokes (9) and bubble Reynolds numbers
(10). An increase in density of particles and in bubble rising veloc-
ity increases the Stokes number. Thus, the bubbleparticle collision
mechanism will change from interception to inertial. This facili-
tates collisions between bubbles and particles. An increase in bub-
ble rising velocity affects also the uid ow conditions around the
bubble surface. At higher bubble rising velocities potential ow
conditions apply, which are more advantageous for bubbleparti-
cle collision than Stokes ow conditions. This can be seen from
Eqs. (16) and (17).
The ow conditions at the bubble surface and the rate of uid
ow drainage between the particle and the bubble are greatly af-
fected by bubble surface retardation. Even small amounts of sur-
face-active materials in the liquid generally render small bubbles
fully retarded, i.e., the tangential velocity at the interface is equal
to zero, and these bubbles rise roughly according to the terminal
velocity of spherical solid particles (Clift et al., 1978; Malysa et
al., 2005). On the other hand, in a carefully puried liquid, the sur-
face of the bubble is completely mobile and the tangential velocity
at the bubbleliquid interface is equal to the uid ow velocity.
When some surfactants are present, they are often swept from
the upper hemisphere to the rear of the bubble, thus the upper
hemisphere is mobile whilst the lower hemisphere is retarded
(Clift et al., 1978; Leja, 1982; Schulze, 1993). The retarded bubble
surface forces the uid streamlines away from the bubbleliquid
interface and hinders thin lm thinning, as a consequence always
giving lower bubbleparticle collision and attachment efciencies
than mobile bubble surfaces (Schulze, 1992, 1993; Dai et al., 2000).
5. Bubbleparticle attachment
Bubbleparticle attachment processes have been less explored
and modelled than those of bubbleparticle collision. This is why
there is only a very limited number of attachment models available
in the literature. In this section, the bubbleparticle attachment
models which are the most signicant for ne particle otation
are reviewed.
The probability of attachment is generally modelled in terms of
a contact time and an induction time. Particlebubble attachment
will occur when the bubbleparticle contact time is longer than the
induction time (Sutherland, 1948). Another way to model bubble
particle attachment is to use the energy barrier approach, where
for bubbleparticle attachment the kinetic energy of a particle
has to be higher than an energy barrier between the particle and
the bubble surface (Yoon and Mao, 1996).
5.1. Induction time
All the particles which collide with the bubble surface do not
necessarily become attached in otation. Some time is required
for the thinning of the intervening liquid lm between a particle
and bubble, lm rupture and the formation of a stable three-
428 T. Miettinen et al. / Minerals Engineering 23 (2010) 420437
phase-contact line (Nguyen et al., 1997). Thus some particles leave
the bubble surface before attachment occurs. The formation of
three-phase-contact line has often been assumed to be very short
for ne and hydrophobic particles and, since the time of lm rup-
ture is of the order of 10
9
s (Yoon and Luttrell, 1989), the thinning
of the intervening liquid lm is considered to be the most impor-
tant component of induction time.
Experimental results indicate that induction time increases
with increasing particle size and decreasing particle surface hydro-
phobicity (Glembotskii, 1953; Ye and Miller, 1988). Ye and Miller
(1988) measured induction times from less than 1 ms to 100 ms
for coal particles of various surface hydrophobicities. Hewitt
et al. (1993) measured the thinning of the intervening liquid lm
under various electrolyte concentrations on clean and weakly
hydrophobic quartz plates and found that the thinning rate of
the intervening liquid lm increased with increasing electrolyte
concentration. Several authors (Blake and Kitchener, 1972; Yoon
and Yordan, 1991; Yoon and Mao, 1996) have measured the critical
lm thicknesses for lm to rupture, and values in the range of 25
300 nm have been reported.
Experimental studies (Trahar and Warren, 1976; Ye and Miller,
1988; Li et al., 1990; Dai et al., 1999) and theoretical suggestions
(Jowett, 1980) show that induction time varies with particle diam-
eter according to a power function
t
ind
= C
1
d
C
2
p
(45)
Jowett (1980) proposed that the parameter C
2
is xed by the
ow conditions experienced by the particle and is equal to 3/2
for coarse particles and 0 for ne particles, and between 0 and 3/
2 for intermediate particles. Furthermore, the parameter C
2
is inde-
pendent of contact angle, electrolyte concentration and bubble
size. This is in agreement with the experimental results of Ye and
Miller (1988) and Dai et al. (1999). The parameter C
1
is propor-
tional to the viscosity of the uid and the effective lm thickness
at which the bubble rst senses the presence of an approaching
particle and inversely proportional to the density of the particle.
In turn, the parameter C
1
decreases with increasing particle contact
angle and bubble size (Dai et al., 1999).
5.2. Contact time
The contact time is considered to be the time for which a parti-
cle and a bubble are in contact after their collision. Bubbleparticle
collision is followed either by particle rebound from the bubble
surface or particle sliding along the bubble surface. In the case of
particle rebound from the bubble surface, the only component in
the contact time is the impact time. If particle sliding occurs after
the bubbleparticle impact, the contact time is the sum of the im-
pact time and the sliding time. Schulze and Gottschalk (1981a,b,c)
have shown that at collision angles smaller than 30, the bubble
particle interaction is an impact process, the contact time being
14 ms. In the case of collision angles larger than about 30, the
bubbleparticle interaction is a sliding process, the sliding time
being 1020 times longer than the impact time.
Particles with diameters less than about 100 lm only impact
and slide on the bubble surface, for their collision kinetic energy
is too small to distort the bubble surface (Dobby and Finch,
1987). There is no rebound without bubble surface deformation.
Typically, the contact times are very short, about 10 ms or less
(Schulze, 1984).
Assuming potential uid ow around a completely mobile bub-
ble surface, Dobby and Finch (1986) derived a sliding time model.
The sliding time was dened as the time it takes a particle to travel
from the point of collision, at an angle h
C
, to the point where it
leaves the bubble surface, i.e., at an angle h = 90 (see Fig. 4)
t
sl
=
d
p
d
b
2(u
p
u
b
) u
b
d
b
dpd
b
_ _
3
ln tan
h
c
2
_ _
: (46)
where u
p
is the particle sedimentation velocity.
5.3. Surface forces
An intervening liquid lm is formed between a particle and a
bubble when the particle and the bubble are in close proximity.
This lm is normally stable between hydrophilic surfaces, whereas
it is usually unstable between a hydrophobic particle and an air
bubble. The stability of the intervening liquid lm can be under-
stood using surface force concepts.
In colloid stability theory the surface forces are divided into
three independent, additive interactions:
v van der Waals
v electrostatic double layer
v non-DLVO.
The rst two interactions are referred to as DLVO interactions,
following Derjaguin and Landau (1941) and Verwey and Overbeek
(1948). These interactions are well investigated, both theoretically
and experimentally.
5.3.1. van der Waals interactions
van der Waals forces are universal attractive forces between
atoms and molecules. These forces also operate between macro-
scopic objects like bubbles and particles. The interaction between
macroscopic bodies arises from spontaneous electric and magnetic
polarizations, which give rise to a uctuating electromagnetic eld
within the media and the gap between them (Elimelech et al.,
1997).
Two methods are available to calculate the van der Waals inter-
actions: the LondonHamaker microscopic approach (Hamaker,
1937; Israelachvili, 1991) and the macroscopic approach of Lifshits
(Lifshits, 1956; Israelachvili, 1991). Calculations using the Lifshits
approach are complex and require detailed information of the
dielectric responses of the interacting media over a very wide fre-
quency range, which makes it unusable for many systems of prac-
tical interest. Hence, it is more convenient to use the London
Hamaker expression. This expression is based on the assumption
of pair wise additivity of intermolecular forces. For two spheres
of radii R
1
and R
2
, surface to surface distance H, the van der Waals
interaction energy, V
vdw
, at close approach H R
i
is given by Eq.
(48)
V
vdw
(H) =
A
132
R
1
R
2
6H(R
1
R
2
)
: (47)
In this expression the Hamaker constant, A
132
, is for the interac-
tion between the two spheres 1 and 2 that are immersed in the
medium 3. The Hamaker constant, A
132
, can be calculated from
the individual Hamaker constants of the medium in vacuum using
Eqs. (49) and (50). The idea in Eq. (49) is that the interaction be-
tween spheres 1 and 2 and medium 3 disappears when the spheres
1 and 2 contact. At the same time, the cohesion of medium 3 is also
considered
A
132
= A
12
A
33
A
13
A
23
(48)
A
12
~ (A
11
A
22
)
1=2
(49)
Since dispersion forces are electromagnetic in character, they
are subjected to a retardation effect (Elimelech et al., 1997). The -
nite time of propagation causes a reduced correlation between
oscillations in the interacting bodies and thus a smaller interaction.
The retardation effect can be introduced into the Hamaker
T. Miettinen et al. / Minerals Engineering 23 (2010) 420437 429
expression by a simple empirical factor. In the case of spheres, the
retarded van der Waals interaction force is given by Eq. (51), where
k is 100 nm for most materials and b = 5.32 (Gregory, 1981). This
expression only applies when the separation is less than about
10% of the particle radius. However, in many cases the interaction
energy is insignicant at larger distances
V
vdw
(H) =
A
132
R
1
R
2
6H(R
1
R
2
)
1
bH
k
ln 1
k
bH
_ _ _ _
: (50)
For the interaction between a particle and bubble, the Hamaker
constant is negative, which results in a repulsive van der Waals
interaction (Fielden et al., 1996; Mishchuk et al., 2002).
5.3.2. Electrostatic double layer interaction
Air bubbles and mineral particles are charged in aqueous media
for various possible reasons, such as the ionisation of surface
groups, specic adsorption of ions, preferential solubility of surface
ions and several others. In the case of metal oxides such as SiO
2
,
TiO
2
, Fe
2
O
3
and MgO, hydroxylated surface groups can lose or gain
a hydrogen ion depending on pH as shown in Eq. (52)
Si OH

2
Si OH Si O

(51)
The pH where the surface charge of the particles is zero is called
the point of zero charge or pzc. With indifferent electrolytes (such
as NaCl, KCl, and KNO
3
) the point of zero charges are also at the pH
at which the zeta potential is zero. In the case of air bubbles, the
point of zero charge, was shown to fall between pH 2 and 3 by
Brandon et al. (1985) and at pH 4 by Takahashi (2005). If the pH
is below pzc, the surface charge of the particles is positive and if
the pH is above, the surface charge of the particles is negative. In
an electrolyte solution, the distribution of oppositely charged
counter ions is accumulated around the charged particles to bal-
ance the surface charge. These counter ions are subjected to elec-
trostatic attraction tending to move the counter ions close to the
particle/bubble surface and the Brownian motion to diffuse the
counter ions randomly around in the solution. The surface charge
on the particle/bubble and the charge of the counter ions around
the particle/bubble constitute the electrical double layer shown
in Fig. 6. The electrical double layer model (Elimelech et al.,
1997; Hunter, 1981) was developed by Stern and later modied
by Grahame.
When a charged bubble and a charged particle come close to
each other, their diffuse layers overlap and they start to interact.
The electrostatic double layer interaction is mainly dependent on
the surface potential, w
s
, and the extent of the diffuse layer, 1/j.
It is believed that the Stern potential is more relevant in the elec-
trostatic interaction than the surface potential, and because the
Stern potential cannot be measured directly, the zeta potential, f,
is an adequate substitute (Hunter, 1981). The zeta potential is
the potential at the plane of shear and can be measured with the
technique of particle electrophoresis. The zeta potential of micro-
bubbles is in a range of 20 mV and 120 mV at pH values of 2
11 (Takahashi, 2005). For quartz particles, the zeta potentials are
between 0 and 110 mV at pH values of 210 (Nguyen and
Fig. 6. The SternGrahame model of the electrical double layer, showing the adsorption of unhydrated ions at the inner Helmholtz plane and of hydrated ions at the outer
Helmholtz plane, together with the diffuse layer which extends outwards into the bulk solution. The electrokinetic shear plane is thought to lie just outside the outer
Helmholtz plane. Note, only counter ions have been shown for simplicity (Elimelech et al., 1997).
430 T. Miettinen et al. / Minerals Engineering 23 (2010) 420437
Schulze, 2004). The magnitude and the extent of the diffuse layer
can be best seen in Eq. (53) which applies in the case of low surface
potentials
w = w
d
exp(ks) (52)
where w
d
is the potential at the Stern layer, j is the DebyeHckel
parameter and w is the potential at distance s from the Stern layer.
The DebyeHckel parameter is an important parameter which
takes into account the effect of ionic strength, dened in Eq. (54).
Eq. (54) is a general form of the DebyeHckel parameter and Eq.
(55) applies to symmetrical zz electrolytes
j
2
=
e
2
Rn
i
z
2
i
e
m
k
B
T
_ _
=
e
2
N
A
Rc
i
z
2
i
e
m
k
B
T
_ _
(53)
j
2
=
2e
2
N
A
cz
2
e
m
k
B
T
_ _
(54)
where e is the charge of an electron, n
i
is the number concentration
of ions in the bulk, c
i
is the concentration of ions in the bulk, z
i
is the
valence of the ions, N
A
is Avogadros number, e
m
is the permittivity
of the medium, k
B
is the Boltzmann constant and T is the tempera-
ture in Kelvin. Eqs. (53) and (55) show that the magnitude and the
extent of the diffuse layer between a charged particle and a charged
bubble can be decreased when the ionic strength is increased, be-
cause the diffuse layer decreases and a greater proportion of the po-
tential drop occurs across the Stern layer, giving a smaller Stern
potential.
To calculate the electrostatic double layer interaction energy, V
e
,
the constant surface potential approximation (CPA) or constant
surface charge approximation (CCA) is usually assumed between
colloidal particles. The constant surface potential case corresponds
to the maintenance of surface chemical equilibrium during ap-
proach, whereas the constant surface charge case might be ex-
pected when the surfaces have a xed surface charge density
(Gregory, 1989). A compromise between these two extremes is
the linear superposition approximation (LSA). The electrostatic
double layer interaction energies for spheres 1 and 2 in all three
cases are given in Eqs. (56)(58)
V
eCPA
(H) =
2pR
1
R
2
n
i
k
B
T
(R
1
R
2
)k
2
(/
2
1
/
2
2
)
2/
1
/
2
/
2
1
/
2
2
ln
1 exp(kH)
1 exp((kH)
_ _
_ _
ln(1 exp(2kH)) (55)
V
eCCA
(H) =
2pR
1
R
2
n
i
k
B
T
(R
1
R
2
)k
2
(/
2
1
/
2
2
)
2/
1
/
2
/
2
1
/
2
2
ln
1 exp(kH)
1 exp((kH)
_ _
_ _
ln(1 exp(2kH)) (56)
V
eLSA
(H) =
128pR
1
R
2
n
i
k
B
T
(R
1
R
2
)k
2
n
1
n
2
exp(kH): (57)
where n = tanh
zew
o
4k
B
T
_ _
and U =
zew
o
k
B
T
are the reduced surface poten-
tials (Elimelech et al., 1997). These equations are only valid at small
separation, H, and for small values of
1,2
and large values of k R
1,2
.
It is reasonable that the electrostatic double layer interaction
energies are repulsive (positive) with surface potentials of like sign
and attractive (negative) with surface potentials of unlike sign at
large surface separation distances. However, at small surface sepa-
ration distances charge reversal may occur and the electrostatic
double layer interaction energy may be attractive with surface
potentials of like sign but unequal magnitude and repulsive with
surface potentials of opposite sign. The former rises in the case of
the constant surface potential approximation and the latter in
the case of the constant surface charge approximation. Experimen-
tal evidence for charge reversal between an air bubble and a at
mica surface is now available (Pushkarova and Horn, 2005).
5.3.3. Non-DLVO interactions
The most signicant non-DLVO forces which can affect bubble
particle attachment are hydration forces, hydrophobic forces and
steric forces. The presence of solid or gas at the surface signi-
cantly perturbs the structure of water adjacent to the surface.
Overlapping of the perturbed water molecules requires work to
be done by or on the system, which leads to repulsive hydration
forces for hydrophilic surfaces and attractive hydrophobic forces
for hydrophobic surfaces. Steric forces arise from macromolecular
reagents which are adsorbed on particles and bubbles. The steric
interaction is mainly repulsive in otation, and these reagents
are often considered to be depressants (Nguyen and Schulze,
2004).
Empirical hydrophobic attractions are usually described using a
single or double exponential term where one or two decay lengths
are invoked (Ralston et al., 2001). Israelachvili and Pashley (1982)
described the hydrophobic force function, F
h
, by an empirical,
exponential function
F
h
R
1
= K
h
e
HH
0
(58)
where K
h
is the hydrophobic force constant and H
0
is the hydropho-
bic force decay length. Yoon and Mao (1996) used a power law to
describe the hydrophobic interaction energy between a particle
and a bubble in the following equation:
V
h
=
R
1
R
2
6(R
1
R
2
)
K
132
H
: (59)
where K
132
is the hydrophobic interaction parameter. It is worth
noting that the hydrophobic interaction energy has the same form
as the nonretarded van der Waals interaction energy Eq. (48).
Recently, Mishchuk et al. (2002) and Snoswell et al. (2003) have
shown theoretically and experimentally that the presence of very
small bubbles on the surface of a particle can lead to an attraction
that appears to be similar to the action of hydrophobic forces. The
foundation of the effect lies in a change in the magnitude and the
sign of the van der Waals interactions between a particle and a
macrobubble. This interaction is repulsive, but as particle surface
hydrophobicity increases, the number, size and height of nucleated
nanobubbles increase, resulting in an attractive macrobubble-sur-
face nanobubble interaction. The experimental data of Ishida and
Higashitani (2005) for the interaction forces between surfaces of
differing hydrophobicities are consistent with the theory men-
tioned above (Mishchuk et al., 2002; Snoswell et al., 2003). On
the other hand, Ishida and Higashitani (2005) measured a short-
ranged attractive force when the surface nanobubbles were re-
moved. This attractive force was of a longer range than the van
der Waals force, and was described as a genuine hydrophobic
attraction (Ishida and Higashitani, 2005). Recent X-ray reectivity
work has shown that water structure adjacent to a smooth hydro-
phobic surface is perturbed only after one or two molecular layers,
thus the origin of this attraction is still obscure (Mezger et al.,
2006).
5.3.4. Extended DLVO theory
In the classical DLVO theory, the total interaction energy is the
sum of the van der Waals and electrostatic double layer energies
(Derjaguin and Landau, 1941; Verwey and Overbeek, 1948).
V
DLVO
(H) = V
vdw
(H) V
e
(H) (60)
In the extended DLVO theory, the hydrophobic interaction en-
ergy is included (Yoon and Mao, 1996)
V
ext-DLVO
(H) = V
vdw
(H) V
e
(H) V
h
(H) (61)
T. Miettinen et al. / Minerals Engineering 23 (2010) 420437 431
For attachment between a bubble and a hydrophobic negatively
charged particle the van der Waals and the electrostatic double
layer interaction energies are repulsive, whereas the hydrophobic
interaction energy is attractive. In the case shown in Fig. 7, the
repulsion outweighs the attraction causing a potential energy bar-
rier. For bubbleparticle attachment, the particle needs to have
sufcient kinetic energy to overcome this energy barrier. If the par-
ticle overcomes the energy barrier, the formed aggregate will be
held in a primary minimum or the thin lm between the particle
and the bubble ruptures and a three-phase contact line is formed.
Another feature in Fig. 7 is a secondary minimum, which is due to
different distance dependencies of the van der Waals, electrostatic
double layer and hydrophobic interaction energies. The depth of
the secondary minimum is usually small and therefore particles
rarely form stable aggregates with bubbles there. In order to en-
hance bubbleparticle attachment, the height of the energy barrier
has to be decreased. This can be done by decreasing the electro-
static double layer repulsion by increasing the ionic strength, or
increasing the hydrophobic attraction by increasing particle sur-
face hydrophobicity. The van der Waals interaction energy gener-
ally cannot be changed unless the temperature is altered or
when surface nanobubbles are present.
5.4. Attachment models
Examples of various kinetic bubbleparticle attachment and
thermodynamic models are given below.
5.4.1. Dobby and Finch model
In a kinetic approach it is clear that the sliding time decreases
with increasing collision angle. Thus, there is a specic collision an-
gle, h
C,A
, at which the sliding time equals the induction time. This
angle is termed the attachment angle or the adhesion angle. All
particles with collision angles smaller than the attachment angle
will attach to the bubble. So the number of attached particles can
be related to the projected area, dened by the attachment angle.
On the other hand, the number of particles which have collided
can be related to the projected area, dened by the maximum pos-
sible collision angle, h
C,max
. Dobby and Finch (1987) proposed that
the bubbleparticle attachment efciency is the ratio of these two
projected areas
E
ADF
=
p(d
b
sinh
C;A
)
2
4
p(d
b
sinh
C;max
)
2
4
or
E
ADF
=
sin
2
h
C;A
sin
2
h
C;max
: (62)
Eq. (47) can be used to calculate the attachment angle, h
C,A
, for
this attachment angle is the collision angle where sliding time, t
cl
,
equals induction time. The value of the attachment angle relates
both the sliding time and induction times to the bubbleparticle
attachment efciency. h
C,max
was assumed to be 90 by Yoon and
Luttrell (Yoon and Luttrell, 1989) and to the angle of tangency, h
t
,
in Eq. (23) (Dai et al., 1999).
5.4.2. Yoon and Mao model
Yoon and Maos thermodynamic approach to the bubbleparti-
cle attachment efciency (Yoon and Mao, 1996), E
A-YM
, is an expo-
nential function containing the ratio of the potential energy
barrier, E
1
, and the kinetic energy of a particle approaching a bub-
ble, E
k
E
A-YM
= exp
E
1
E
k
_ _
(63)
The average kinetic energy of particles was calculated from the
radial velocity of an approaching particle using Yoon and Luttrells
empirical stream function Eq. (13). The maximum value of the po-
tential energy barrier is determined from the extended DLVO the-
ory, where the total interaction energy is a sum of the electrostatic
double layer, V
e
, van der Waals, V
vdw
, and hydrophobic, V
h
, interac-
tion energies.
Eq. (63) suggests that if the kinetic energy of the particle, E
k
, is
much greater than E
1
, the thin intervening lm will rupture and a
three-phase contact line will form between a bubble and a particle.
It can also be seen that the bubbleparticle attachment efciency
increases if E
1
decreases, which can be achieved by increasing
the hydrophobicity of particles and/or reducing the electrostatic
repulsion between negatively charged particles and bubbles. The
attachment efciency also increases with increasing E
k
, which
may be achieved by providing high-shear agitation.
Due to insufcient information available for the hydrophobic
forces for bubbleparticle interactions, Yoon and Mao (1996)
back-calculated the K
132
values (Eq. (60)) from the otation rate
constants, determined using the data from single bubble otation
experiments. The K
132
values were in the range of 0.61.7
10
19
J. Attachment efciencies calculated with the Yoon and
Mao model were found to increase with particle size as a result
of an overestimation of the kinetic energy of the larger particles,
which is in contradiction with common knowledge (Dai et al.,
1999; Mishchuk et al., 2001).
5.4.3. Scheludko model
The only thermodynamic approach for the limit of otation of
ne particles has been performed by Scheludko et al. (1976). They
stated that the kinetic energy of ne particles has to be larger than
the energy needed to disrupt the intervening liquid lm and form a
three-phase contact line if bubbleparticle attachment is to occur.
When these energies are in balance, an equation for the minimum
particle diameter, for particle otation d
p(min)
, can be derived, for a
given receding contact angle h
r
d
p(min)
= 2
3L
2
u
2
b
c
LV
(q
p
q
f
)(1 cos h
r
)
_ _
1=3
(64)
0 1 100 H (nm)
V

(
k
T
)
Electrostatic
double layer
interaction
Van der Waals
interaction
Hydrophobic
interaction
Energy
barrier
Primary
minimum
Secondary
minimum
Extended DLVO
interaction
0
Fig. 7. Schematic potential energy diagram for the interaction of a bubble and a
hydrophobic particle. The electrostatic double layer repulsion, V
e
(H), the van der
Waals repulsion, V
vdw
(H), the hydrophobic attraction V
h
(H) and the total extended
DLVO energy, V
ext-DLVO
(H), are shown as a function of surface-to-surface distance, H.
432 T. Miettinen et al. / Minerals Engineering 23 (2010) 420437
where L is the solidliquidvapour three-phase contact line tension
and c
LV
is the liquidvapour interfacial tension. The contact line it-
self introduces a contribution to the free energy of the solidliquid
vapour system, as rst proposed by Gibbs (1928). The line tension is
added to Youngs equation (Young, 1805):
cos h
L
=
c
SV
c
SL
c
LV

L
R
L
c
LV
= cos h
Y

L
R
L
c
LV
(65)
where h
L
is the equilibrium contact angle, including line tension, c
SL
is the solidliquid interfacial tension, c
SV
is the solidvapour inter-
facial tension, h
Y
is the Young contact angle and R
L
is the radius of
curvature of the contact line in the plane of solid. For positive values
of line tension, the line tension will tend to contract the contact line
and increase the contact angle; the reverse situation holds for neg-
ative values of line tension. The magnitude and the sign of the line
tension have been debated for many years (Widom, 1995; Solo-
montsev and White, 1999; Wang et al., 1999; Pompe and Herming-
haus, 2000; Pompe, 2002; Checco et al., 2003). In Eq. (64) the sign of
the line tension is not important because the value is squared. The
magnitude is critical, however. Scheludko et al. (1976) estimated
the magnitude of the line tension to be between 2.8 and
5.6 10
10
N from Eq. (64) under conditions that normally occur
in otation (u
b
= 0.2 m/s; c
LV
= 50 mN/m; q
p
= 3000 kg/m
3
; q
f
=
1000 kg/m
3
) for 2 lm particles in diameter with contact angles
between 20 and 40. Recently, Yang et al. (2003) studied the for-
mation of nanobubbles at structured solidwater interfaces using
the tapping mode atomic force microscopy (TMAFM) imaging tech-
nique. The difference between microscopic contact angles detected
by TMAFM and macroscopic contact angles was linked to the inu-
ence of the line tension. The calculated line tension of the bubbles,
obtained by imaging the bubble from above so that the local radius
of curvature could be obtained, was 3.0 10
10
N. This value is
similar in magnitude to the reported values of line tension of small
droplets (Pompe and Herminghaus, 2000), the values determined
by Scheludko et al. (1976) and as expected theoretically (Drelich,
1996). There is good agreement between the limits of otation for
particles around 1 lm in diameter and the Scheludko model (Gon-
tijo et al., 2007) despite real concerns about its conceptual base, i.e.
how does such a small particle acquire sufcient kinetic energy to
disrupt the intervening lm?
5.5. Factors inuencing bubbleparticle attachment
The factors inuencing bubbleparticle attachment are re-
viewed from both thermodynamic and kinetic points of view.
Thermodynamic analyses show that the interaction energy
barrier between a particle and a bubble decreases with increasing
particle surface hydrophobicity independently of whether hydro-
phobic force (Yoon and Mao, 1996) or surface nanobubble (Mish-
chuk et al., 2002; Snoswell et al., 2003) approaches are used. This
has also been experimentally veried (Ducker et al., 1994; Fielden
et al., 1996; Ishida and Higashitani, 2005). From a kinetic point of
view, theoretical studies show that an increase in particle surface
hydrophobicity results in earlier rupture of the aqueous lm be-
tween a particle and bubble. Film drainage rate measurements
by Hewitt et al. (1993) indicated that a higher drainage rate is ob-
tained for more hydrophobic surfaces. The experimental data of
Newcombe and Ralston (1994) showed that the three-phase con-
tact line expansion rate increased with increasing surface hydro-
phobicity. Ye and Millers (1988) measurements showed shorter
induction times for more hydrophobic coal particles. Thus, an in-
crease in particle surface hydrophobicity is benecial to bubble
particle attachment.
It has been extensively reported in the literature that the exper-
imental otation rates and recoveries increase with increasing
electrolyte concentration (Klassen and Mokrousov, 1963; Collins
and Jameson, 1977; Crawford and Ralston, 1988; Dai et al.,
1999). The positive effect of electrolyte concentration on the ota-
tion rates and recoveries is mainly attributed to an increase in the
bubbleparticle attachment efciency. From a thermodynamic
point of view, this is explained by a decreased electrostatic double
layer interaction. The values of both the particle (Yoon and Mao,
1996; Snoswell et al., 2003) and bubble (Li and Somasundaran,
1991; Takahashi, 2005) zeta potentials decrease with increasing
electrolyte concentration. In turn, the lm drainage rate measured
by Hewitt et al. (1993) increased with increasing ionic strength,
whereas the critical lm thickness measured by Blake and Kitchen-
er (1972) decreased with increasing ionic strength. In addition to
the kinetic and thermodynamic inuences of electrolyte concen-
tration on bubbleparticle attachment efciency, another possible
reason for the increased bubbleparticle attachment efciency
with electrolyte concentration is based on a decrease in gas solu-
bility as electrolyte concentration increases (Weisenberger and
Schumpe, 1996). This will result in the precipitation of gas mole-
cules as nanobubbles on the particle surface, which is advanta-
geous for bubbleparticle attachment (Dai et al., 1998).
The inuence of pH on bubbleparticle attachment depends on
the zeta potentials of bubbles and particles. Usually in otation,
both the surfaces are negatively charged and an increase in pH
leads to an increase in the values of their zeta potentials (Laskow-
ski and Kitchener, 1969; Brandon et al., 1985; Kelsall et al., 1996;
Snoswell et al., 2003; Takahashi, 2005). This causes an increase
in electrostatic double layer repulsion between these surfaces
and a decrease in the bubbleparticle attachment efciency. There
is also experimental evidence that particle contact angle decreases
as the pH decreases below or increases above the pzc (Laskowski
and Kitchener, 1969).
According to the experimental data of Hewitt et al. (1995) and
Dai et al. (1999), bubbleparticle attachment efciency increases
with decreasing bubble and particle size. Hewitt et al. (1993) ex-
plained the effect of bubble size by means of the shorter lm drain-
age time associated with smaller bubbles. Newcombe and Ralston
(1994) also showed experimentally that smaller bubbles have
shorter three-phase contact line expansion times than do large
bubbles. The inuence of particle size can be explained in terms
of the decrease in sliding time and the increase in induction time
with increasing particle size. Particle shape has also a strong effect
on the bubbleparticle attachment efciency. Many researchers
(Anfruns and Kitchener, 1977; Blake and Ralston, 1985; Dai,
1998) have also found that angular particles have higher otation
recovery or collection efciency than smooth spheres. Anfruns and
Kitchener (1977) explained the higher bubbleparticle collection
efciency of angular particles by means of the facilitated lm rup-
ture between a particle and a bubble due to asperities on the par-
ticle surface.
6. Bubbleparticle stability
When the three-phase contact line between a bubble and a par-
ticle is formed, it is able to resist strong detachment forces. These
detachment forces involve inertia, gravity and viscous forces and
increase with increasing particle size, for the mass of the particle
and the area exposed to the detaching uid ow increases. Derja-
guin et al. (1984) estimated that the detachment forces are a mil-
lion times greater when the particle size is about 100 lm than if
the particle size is about 1 lm. Thus, it is often assumed that the
bubbleparticle stability efciency is very close to unity for parti-
cles less than 1 lm, especially if they are strongly hydrophobic.
In this section, the physics behind the bubbleparticle stability
models under quiescent and turbulent conditions are presented.
T. Miettinen et al. / Minerals Engineering 23 (2010) 420437 433
An example of calculation of the bubbleparticle stability ef-
ciency under turbulent conditions is also shown as a function of
particle size, particle hydrophobicity and energy dissipation in a
otation cell.
6.1. Stability model
Under quiescent conditions, the forces acting on a bubblepar-
ticle aggregate include the capillary force, F
ca
, the gravity of the
particle, F
g
, the hydrostatic pressure of the liquid, F
hyd
, and the sta-
tic buoyancy of the immersed part of the particle by the liquid, F
b
(Sutherland and Wark, 1955). This model corresponds to the con-
ditions experienced when a captive bubble is gently pressed
against a given particle (Crawford and Ralston, 1988). A more real-
istic bubbleparticle stability model under turbulent conditions
was proposed by Schulze (1984).
6.1.1. Schulze model
According to Schulze (1984), the bubbleparticle stability ef-
ciency, E
S
, can be assumed to be exponentially distributed and ex-
pressed by
E
S
= 1 exp 1
1
Bo
_ _
(66)
where Bo
/
is the Bond number. The Bond number describes the sta-
bility of the bubbleparticle aggregate and is characterized by the
ratio of the detachment forces to the attachment forces
Bo
/
=
F
g
F
b
F
d
F
c
F
ca
F
hyd
(67)
where F
d
is the additional detaching forces and F
c
is the capillary
pressure in the gas bubble.
The expressions for the various forces in Eq. (67) are:
F
g
=
4
3
pR
3
p
q
p
g (68)
where g is the gravitational acceleration. The static buoyancy of the
immersed part of the particle by the liquid is
F
b
=
p
3
R
3
p
q
f
g[(1 cos x)
2
(2 cos x)[: (69)
where x is the centre-angle between the rear part of the attached
sphere and three-phase contact line projection area on the sphere
as shown in Fig. 8.
The force for hydrostatic pressure of the liquid above the three-
phase contact line area
F
hyd
= pr
2
0
q
f
gz
0
= pR
2
p
(sin
2
x)q
f
gz
0
: (70)
where r
0
is the three-phase contact radius and z
0
is the liquid height
as shown in Fig. 8.
The capillary force on the three-phase contact with radius r
0
is
F
ca
= 2pc
LV
R
p
sinxsin(xh) (71)
where h is the particle contact angle as shown in Fig. 8.
The additional detaching forces can be presented as the product
of the particle mass and the acceleration a in the external eld of
ow
F
d
=
4
3
pR
3
p
q
p
a (72)
The capillary pressure in the gas bubble, which acts on the con-
tact area of the attached particle, can be expressed as
F
r
= pR
2
p
(sin
2
x)
2c
LV
R
b
2R
b
q
f
g
_ _
(73)
In Eq. (72) the detachment forces and the acceleration a in the
external eld of ow can be related to the turbulent ow eld
and the dissipation energy, e. Schulze assumed that aggregates
are moved mainly by the centrifugal acceleration present in the
vortex and the turbulent vortex radius is equal to the aggregate ra-
dius. Thus the acceleration a can be described as
a ~ 1:9
e
2=3
d
b
2

dp
2
_ _
1=3
(74)
In otation machines the mean energy dissipation is in the
range from 1 to 100 W/kg. Accordingly, the acceleration involved
is of the order of 2200 g units.
Fig. 8. The gassolidliquid three-phase contact for a smooth spherical particle. See text for explanation Schulze (1984). Reproduced with permission from Elsevier, Schulze,
H.J., Physico-chemical elementary processes in otation: an analysis from the point of view of colloid science including process engineering considerations, Elsevier,
Amsterdam, New York (1984).
434 T. Miettinen et al. / Minerals Engineering 23 (2010) 420437
The force balance can only be calculated using the numerical
integration of the Laplace equation of capillary meniscus. The La-
place equation gives the meniscus deformation z
0
as a function
of particle size. An example of calculation of the bubbleparticle
stability as a function of particle radius, particle contact angle
and energy dissipation is shown in Fig. 9. The calculations in
Fig. 9 show that aggregates between bubbles and ne particles
are very stable. If the bubbleparticle attachment models, e.g.,
the Scheludko model above, are grossly compared with Fig. 9, it
can be seen that if a 1 lm particle is able to form a three-phase
contact line with a bubble, the aggregate formed is stable even in
turbulent conditions.
7. Summary
In the future, many otation operations need to improve liber-
ation by grinding minerals to ner sizes, in order to increase recov-
ery for low grade and nely disseminated mineral deposits.
The low otation rate and recovery of hydrophobic ne parti-
cles (<20 lm) is mainly due to their low collision efciency, E
C
,
with bubbles. Their E
C
values can be increased by decreasing the
bubble size and by aggregating the ne particles to an optimum
size for otation.
A decrease in the bubble size not only increases the bubblepar-
ticle collision efciency but also increases the bubbleparticle
attachment efciency and the number of generated bubbles in
the case of constant gas ow rate. These factors also increase the
otation rate and recovery of ne particles but might cause higher
water recovery, which increases entrainment of gangue minerals.
The bubble size can be decreased by mechanical and physicochem-
ical methods. In mineral otation mechanical methods are more
common, whereas physicochemical methods have been widely
used in water treatment practice, where selectivity is not needed.
For ne particles, it is experimentally and theoretically clear
that the otation rate increases with increasing particle size. Thus,
many techniques have been developed which try to increase parti-
cle size and mass and decrease surface energy. All these techniques
have the same feature that ne particles are induced to form ocs
or aggregates. Again, a lack of selectivity in aggregation has re-
stricted their applicability in mineral otation.
Various bubbleparticle collision and attachment efciency
models have been reviewed and the factors inuencing collision
and attachment efciencies were discussed. It was seen that for
ne particles, the main collision mechanism is interception,
whereas submicron particles are also affected by Brownian motion,
and larger particles by inertia. Bubbleparticle attachment models
were developed on the premise that the particle sliding time has to
be longer than the induction time or the kinetic energy of the par-
ticle has to be larger than the energy barrier between the bubble
and the particle. Both theoretical and experimental studies in the
literature showed that the bubbleparticle attachment efciencies
in potential and Stokes ow conditions increase with decreasing
particle and bubble size and increasing particle contact angle and
electrolyte concentration.
For ne particle otation, the bubbleparticle stability can be
assumed to equal unity because if ne particles are able to form
a three-phase contact line with a bubble, the formed aggregate is
stable even in turbulent conditions.
In practice, ne particle otation can be improved by allowing
long residence times and working at high collector coverages (large
contact angles). New approaches are required, which could include
very high energy zones for bubbleparticle contact or completely
novel ways of introducing particles directly to the watervapour
interface.
Acknowledgements
Financial support from the Australian Research Council and
AMIRA International is gratefully acknowledged.
References
Abrahamson, J., 1975. Collision rates of small particles in a vigorously turbulent
uid. Chemical Engineering Science 30 (11), 13711379.
Acheson, D.J., 1990. Elementary Fluid Dynamics. Oxford University Press, New York.
p. 408.
Ahmed, N., Jameson, G.J., 1985. The effect of bubble size on the rate of otation of
ne particles. International Journal of Mineral Processing 14 (3), 195215.
Anfruns, J.F., Kitchener, J.A., 1977. Rate of capture of small particles in otation.
Transactions of the Institution of Mining and Metallurgy, Section C: Mineral
Processing and Extractive Metallurgy 86, 915.
Attia, Y.A., 1977. Synthesis of PAMG chelating polymers for the selective
occulation of copper minerals. International Journal of Mineral Processing 4
(3), 191208.
Attia, Y.A., 1982. Fine particle separation by selective occulation. Separation
Science and Technology 17 (3), 485493.
Batchelor, G.K., 2000. An Introduction to Fluid Dynamics. Cambridge University
Press, Cambridge. p. 635.
Bennett, A.J.R., Chapman, W.R., Dell, C.C., 1958. Froth otation of coal. International
Coal Preparation Congress, vol. E2, third ed. Brussels-Liege, pp. 452462.
Blake, P., Ralston, J., 1985. Controlled methylation of quartz particles. Colloids and
Surfaces 15 (1-21-2), 101118.
Blake, T.D., Kitchener, J.A., 1972. Stability of aqueous lms on hydrophobic
methylated silica. Journal of the Chemical Society, Faraday Transactions 1:
Physical Chemistry in Condensed Phases 68 (pt. 8), 14351442.
Brandon, N.P., Kelsall, G.H., Levine, S., Smith, A.L., 1985. Interfacial electrical
properties of electrogenerated bubbles. Journal of Applied Electrochemistry 5
(4), 485493.
Checco, A., Guenoun, P., Daillant, J., 2003. Nonlinear dependence of the contact
angle of nanodroplets on contact line curvature. Physical Review Letters 91
(18), 186101186104.
Clift, R., Grace, J.R., Weber, M.E., 1978. Bubbles Drops and Particles. Academic Press,
New York. p. 400.
Collins, G.L., Jameson, G.J., 1977. Double-layer effects in the otation of ne
particles. Chemical Engineering Science 32 (3), 239246.
Collins, G.L., 1975. Dispersed air otation of ne particles. PhD Thesis, University of
London.
Fig. 9. Bubbleparticle stability efciency, E
s
, as a function of particle radius, R
p
, for
different values of contact angle, h and energy dissipation, e R
p
(q
p
= 2500 kg/m
3
;
q
f
= 1000 kg/m
3
; c
LV
= 70 mN/m; R
b
= 0.5 mm) (Schulze, 1984). Reproduced with
permission from Elsevier, Schulze, H.J., Physico-chemical elementary processes in
otation: an analysis from the point of view of colloid science including process
engineering considerations, Elsevier, Amsterdam, New York (1984).
T. Miettinen et al. / Minerals Engineering 23 (2010) 420437 435
Crawford, R., Ralston, J., 1988. The inuence of particle size and contact angle in
mineral otation. International Journal of Mineral Processing 23 (12), 124.
Dai, Z., 1998. Particlebubble heterocoagulation. PhD Thesis, University of South
Australia.
Dai, Z., Fornasiero, D., Ralston, J., 1998. Inuence of dissolved gas on bubbleparticle
heterocoagulation. Journal of the Chemical Society Faraday Transactions 94
(14), 19831987.
Dai, Z., Dukhin, S.S., Fornasiero, D., Ralston, J., 1998. The inertial hydrodynamic
interaction of particles and rising bubbles with mobile surfaces. Journal of
Colloid and Interface Science 197 (2), 275292.
Dai, Z., Fornasiero, D., Ralston, J., 1999. Particlebubble attachment in mineral
otation. Journal of Colloid and Interface Science 217 (1), 7076.
Dai, Z., Fornasiero, D., Ralston, J., 2000. Particlebubble collision models a review.
Advances in Colloid and Interface Science 85 (23), 231256.
Davis, R.E., Acrivos, A., 1966. Inuence of surfactants on the creeping motion of
bubbles. Chemical Engineering Science 21 (8), 681685.
Derjaguin, B.V., Landau, L., 1941. Theory of the stability of strongly charged
lyophobic sols and of the adhesion of strongly charged particles in solutions of
electrolytes. Acta Phys. Chim. 14, 633.
Derjaguin, B.V., Dukhin, S.S., 1961. Theory of otation of small and medium-size
particles. Bulletin Institution of Mining and Metallurgy 651, 21246.
Derjaguin, B.V., Dukhin, S.S., Rulyov, N.N., 1984. Kinetic theory of otation of small
particles. Surface and Colloid Science 13, 71113.
Dobby, G.S., Finch, J.A., 1986. A model of particle sliding time for otation size
bubbles. Journal of Colloid and Interface Science 109 (2), 493498.
Dobby, G.S., Finch, J.A., 1987. Particle size dependence in otation derived from a
fundamental model of the capture process. International Journal of Mineral
Processing 21 (34), 241260.
Drelich, J., 1996. The signicance and magnitude of the line tension in three-phase
(solidliquiduid) systems. Colloids and Surfaces A: Physicochemical and
Engineering Aspects 116 (1/2), 4354.
Duan, J., Fornasiero, D., Ralston, J., 2003. Calculation of the otation rate constant of
chalcopyrite particles in an ore. International Journal of Mineral Processing 72
(14), 227237.
Ducker, W.A., Xu, Z., Israelachvili, J.N., 1994. Measurements of hydrophobic and
DLVO forces in bubblesurface interactions in aqueous solution. Langmuir 10
(9), 32793289.
Dziensiewicz, J., Pryor, E.J., 1950. An investigation into the action of air in froth
otation. Bulletin Institution of Mining and Metallurgy 521, 122.
Edwards, D.A., Brenner, H., Wasan, D.T., 1991. Interfacial Transport Processes and
Rheology. Butterworth-Heinemann, Boston. p. 558.
Elimelech, M., Williams, R.A., Jia, X., Gregory, J., 1997. Particle Deposition and
Aggregation: Measurement, Modelling and Simulation. Butterworth-
Heinemann, Woborn, MA. p. 448.
Fielden, M.L., Hayes, R.A., Ralston, J., 1996. Surface and capillary forces affecting air
bubbleparticle interactions in aqueous electrolyte. Langmuir 12 (15), 3721
3727.
Fuerstenau, D.W., 1980. Fine particle otation. In: Fine Particle Processing,
Proceedings International Symposium, vol. 1, pp. 669705.
Gaudin, A.M., Schuhmann Jr., R., Schlechten, A.W., 1942. Flotation kinetics II. The
effect of size on the behaviour of galena particles. Journal of Physical Chemistry
46, 902910.
Gaudin, A.M., 1957. Flotation, second ed. McGraw-Hill Book Company, New York.
560.
Gibbs, J.W., 1928. The Collected Works of J. Willard Gibbs, vol. 1. Longmans-Green,
New York.
Glembotskii, V.A., 1953. Rate of adhesion of air-bubbles to mineral particles during
otation, and themods for its measurement. Izvestiya Akademii Nauk SSSR,
Otdelenie Tekhnicheskikh Nauk, 15241531.
Glembotskii, V.A., Mamakov, A.A., Romanov, A.M., Nenno, V.E., 1975. Selective
separation of ne mineral slimes using the method of electric otation. In:
Proceedings 11th International Mineral Processing Congress, pp. 561582.
Gontijo, C., Fornasiero, D., Ralston, J., 2007. The limits of coarse particle otation.
Canadian Journal of Chemical Engineering, 739747 (special issue, October).
Gorain, B.K., Franzidis, J.-P., Manlapig, E., 1997. Studies on impeller type, impeller
speed and air ow rate in an industrial scale otation cell. Part 4. Effect of
bubble surface area ux on otation performance. Minerals Engineering 10 (4),
367379.
Gregory, J., 1981. Approximate expressions for retarded van der Waals interaction.
Journal of Colloid and Interface Science 83 (1), 138145.
Gregory, J., 1989. Fundamentals of occulation. Critical reviews in Environmental
Control 19 (3), 185230.
Hamaker, H.C., 1937. The Londonvan der Waals attraction between spherical
particles. Physica (The Hague) 4, 10581072.
Hewitt, D., Fornasiero, D., Ralston, J., Fisher, L.R., 1993. Aqueous lm drainage at the
quartz/water/air interface. Journal of the Chemical Society, Faraday
Transactions 89 (5), 817822.
Hewitt, D., Fornasiero, D., Ralston, J., 1995. Bubbleparticle attachment. Journal of
the Chemical Society, Faraday Transactions 91 (13), 19972001.
Hoover, R.M., Malhotra, D., 1976. Emulsion otation of molybdenite. Flotation 1,
485505.
Hunter, R.J., 1981. Colloid Science: Zeta Potential in Colloid Science: Principles and
Applications. Academic Press, London. p. 386.
Ishida, N., Higashitani, K., 2005. Interaction forces between hydrophobic surfaces
evaluated by AFM the role of nanoscopic bubbles in the interactions.
Publications of the Australasian Institute of Mining and Metallurgy (Centenary
of Flotation Symposium), pp. 481486.
Israelachvili, J.N., Pashley, R., 1982. The long-range hydrophobic interaction
decaying exponentially with distance. Nature (London, United Kingdom) 300
(5890), 341.
Israelachvili, J.N., 1991. Intermolecular and Surface Forces, second ed. Academic
Press, San Diego. p. 291.
Jameson, G.J., Nam, S., Young, M.M., 1977. Physical factors affecting recovery rates
in otation. Minerals Science Engineering 9 (3), 103118.
Johnson, R.W., 1998. The Handbook of Fluid Dynamics. CRC Press, Boca Raton. p.
1952.
Jordan, C.E., Spears, D.R., 1990. Evaluation of a turbulent ow model for ne-bubble
and ne-particle otation. Minerals and Metallurgical Processing 7 (2), 6573.
Jowett, A., 1980. Formation and disruption of particlebubble aggregates in
otation. In: Fine Particle Processing, Proceedings International Symposium,
vol. 1, pp. 720754.
Kelsall, G.H., Tang, S., Yurdakul, S., Smith, A.L., 1996. Electrophoretic behaviour of
bubbles in aqueous electrolytes. Journal of the Chemical Society, Faraday
Transactions 92 (20), 38873893.
Klassen, V.I., Mokrousov, V.A., 1963. An introduction to the theory of otation.
Butterworths, London. p. 493.
Koh, P.T.L., Warren, L.J., 1980. A pilot plant test of the shear-occulation of ultrane
scheelite. In: Chemeca 80 (Eighty): Process Industries: 80s; 8th Australian
Chemical Engineering Conference, pp. 9094.
Lai, R.W.M., Fuerstenau, D.W., 1968. Liquidliquid extraction of ultrane particles.
Transactions of the Society of Mining Engineers of AIME 241 (4), 549556.
Laskowski, J., Kitchener, J.A., 1969. Hydrophilichydrophobic transition on silica.
Journal of Colloid and Interface Science 29 (4), 670679.
Leja, J., 1982. Surface Chemistry of Froth Flotation. Plenum Press, New York. p. 758.
Levich, V.G., 1962. Physicochemical Hydrodynamics. Prentice-Hall, Englewood
Cliffs, NJ, USA. p. 700.
Li, C., Somasundaran, P., 1991. Reversal of bubble charge in multivalent inorganic
salt solutions effect of magnesium. Journal of Colloid and Interface Science
146 (1), 215218.
Li, D., Fitzpatrick, J.A., Slattery, J.C., 1990. Rate of collection of particles by otation.
Industrial and Engineering Chemistry Research 29 (6), 955967.
Lifshits, E.M., 1956. The theory of molecular attraction forces between solid bodies.
Soviet Physics, JETP 2, 7383.
Liu, Q., Wannas, D., 2004. The role of polymeric-depressant-induced occulation in
ne particle otation. In: Particle Size Enlargement in Mineral Processing,
Proceedings of the UBC-McGill Biennial International Symposium on
Fundamentals of Mineral Processing 5th, Hamilton, Canada, 225 August, pp.
179193.
Malysa, K., Krasowska, M., Krzan, M., 2005. Inuence of surface active substances on
bubble motion and collision with various interfaces. Advances in Colloid and
Interface Science, 205225.
Matis, K.A., Backhurst, J.R., 1984. Laboratory studies of electrolytic otation as a
separation technique. SolidLiq Sep. (ap.-Symp. Adv. SolidLiq. Sep.), 2940.
Matis, K.A., Gallios, G.P., 1986. Dissolved air and electrolytic otation. NATO ASI
Series, Series E: Applied Sciences 117, 3769 (Miner. Process. Crossroads).
Matis, K.A., Gallios, G.P., Kydros, K.A., 1993. Separation of nes by otation
techniques. Separation Technology 3 (2), 7690.
Matis, K.A., 1995. Flotation Science and Engineering. Marcel Dekker, New York.
Matis, K.A., Zouboulis, A.I., 1995. Electrolytic otation: an unconventional
technique. Flotation Science and Engineering, 385413.
Mezger, M., Reichert, R., Schoder, S., Okasinski, J., Schroder, H., Dosch, H., Palms, D.,
Ralston, J., Honkimaki, V., 2006. The water gap at hydrophobic interfaces. In:
Proceedings National Academy Science, vol. 103(49), pp. 1840118404.
Mishchuk, N., Koopal, L.K., Dukhin, S.S., 2001. Microotation suppression and
enhancement caused by particle/bubble electrostatic interaction. Journal of
Colloid and Interface Science 237 (2), 208223.
Mishchuk, N., Ralston, J., Fornasiero, D., 2002. Inuence of dissolved gas on van der
Waals forces between bubbles and particles. Journal of Physical Chemistry A
106 (4), 689696.
Mishchuk, N., Ralston, J., Fornasiero, D., 2006. Inuence of very small bubbles on
particle/bubble heterocoagulation. Journal of Colloid and Interface Science 301,
168177.
Newcombe, G., Ralston, J., 1994. Bubble spreading kinetics and mineral otation.
Minerals Engineering 7 (7), 889903.
Nguyen, A.V., Schulze, H.J., Ralston, J., 1997. Elementary steps in particlebubble
attachment. International Journal of Mineral Processing 51 (14), 183195.
Nguyen, A.V., Schulze, H.J., 2004. Colloidal Science of Flotation, vol. 118. Marcel
Dekker, New York. p. 850.
Nguyen, A.V., George, P., Jameson, G.J., 2006. Demonstration of a minimum in the
recovery of nanoparticles by otation: theory and experiment. Chemical
Engineering Science 61 (8), 24942509.
Peng, Y., Cotnoir, D., Ourriban, M., Richard, D., Liu, Q., 2005. Some solutions to the
problems in ne particle otation. In: Centenary of Flotation Symposium.
Publications of the Australasian Institute of Mining and Metallurgy, pp. 535
540.
Pompe, T., Herminghaus, S., 2000. Three-phase contact line energetic from
nanoscale liquid surface topographies. Physical Review Letters 85 (9), 1930
1933.
Pompe, T., 2002. Line tension behaviour of a rst-order wetting system. Physical
Review Letters 89 (7), 076102.
436 T. Miettinen et al. / Minerals Engineering 23 (2010) 420437
Prieve, D.C., Ruckenstein, E., 1974. Effect of London forces upon the rate of
deposition of Brownian particles. AIChE Journal 20 (6), 11781187.
Pushkarova, R.A., Horn, R.G., 2005. Surface forces measured between an air bubble
and a solid surface in water. Colloids and Surfaces A: Physicochemical and
Engineering Aspects 261 (13), 147152.
Pyke, B.L., Fornasiero, D., Ralston, J., 2003. Bubble particle heterocoagulation under
turbulent conditions. Journal of Colloid and Interface Science 265 (1), 141151.
Raju, G.B., Khangaonkar, P.R., 1982. Electrootation of chalcopyrite nes.
International Journal of Mineral Processing 9 (2), 133143.
Raju, G.B., Khangaonkar, P.R., 1984. Electrootation of chalcopyrite nes with
sodium diethyldithiocarbamate as collector. International Journal of Mineral
Processing 13 (3), 211221.
Ralston, J., 1992. The inuence of particle size and contact angle in otation.
Developments in Minerals Processing 12, 203224 (Colloid Chemistry Mineral
Processing).
Ralston, J., Dukhin, S.S., 1999. The interaction between particles and bubbles.
Colloids and Surfaces A: Physicochemical and Engineering Aspects 151 (12),
314.
Ralston, J., Fornasiero, D., Mishchuk, N., 2001. The hydrophobic force in otation a
critique. Colloids and Surfaces A: Physicochemical and Engineering Aspects 192
(13), 3951.
Ralston, J., Dukhin, S.S., Mishchuk, N.N., 2002. Wetting lm stability and otation
kinetics. Advances in Colloid and Interface Science 95, 145236 (2,3).
Rao, G.V., 1997. Spherical agglomeration of scheelite nes with amphoteric
collector. Metals, Materials and Processes 9 (1), 5763.
Reay, D., Ratcliff, G.A., 1973. Removal of ne particles from water by dispersed air
otation. Effects of bubble size and particle size on collection efciency..
Canadian Journal of Chemical Engineering 51, 178185.
Reay, D., Ratcliff, G.A., 1975. Experimental testing of the hydrodynamic collision
model of ne particle otation. Canadian Journal of Chemical Engineering 53
(5), 481486.
Romanov, A.M., 1998. Electrootation in wastewater treatment: results and
perspectives. NATO ASI Series, Series 2: Environment 43, 335360 (Mineral
Processing and the Environment).
Rubio, J., Capponi, F., Matiolo, E., Nunes, G.N., 2003. Advances in otation of mineral
nes. In: XXII International Mineral Processing Congress, pp. 10141022.
Rulyov, N.N., 1989. Colloidal-hydrodynamic otation theory. Khimiya i Teknologiya
Vody 11 (3), 195216.
Rulyov, N.N., 2001. Turbulent microotation: theory and experiment. Colloids and
Surfaces A: Physicochemical and Engineering Aspects 192 (13), 7391.
Scheludko, A., Toshev, B.V., Bojadjiev, D.T., 1976. Attachment of particles of a uid
surface (capillary theory of otation). Journal of the Chemical Society, Faraday
Transactions 1: Physical Chemistry in Condensed Phases 72 (12), 28152828.
Schwarz, S., Grano, S.R., 2002. Effect of particle hydrophobicity on particle and
water transport across a otation froth. In: Flotation and Flocculation: From
Fundamentals to Applications. University of South Australia, Kailua-Kona,
Hawaii.
Schulze, H.J., Gottschalk, G., 1981a. Experimental studies of the hydrodynamic
interaction of particles with a gas bubble. Kolloidnyi Zhurnal 43 (5), 934
944.
Schulze, H.J., Gottschalk, G., 1981b. Investigations of the hydrodynamic interaction
between a gas bubble and mineral particles in otation. Developments in
Mineral Processing 2, 6385 (Mineral Processing, Part A).
Schulze, H.J., Gottschalk, G., 1981c. Preliminary experimental investigation of
hydrodynamic interaction between particles and a gas bubble. Aufbereitungs
Technik (19601989) 22 (5), 254264.
Schulze, H.J., 1984. Physico-chemical Elementary Processes in Flotation: An Analysis
from the Point of View of Colloid Science Including Process Engineering
Considerations. Elsevier, Amsterdam, New York.
Schulze, H.J., 1992. Probability of particle attachment on gas bubbles by sliding.
Advances in Colloid and Interface Science 40, 82838305.
Schulze, H.J., 1993. Flotation as a heterocoagulation process: possibilities of
calculating the probability of otation. Surfactant Science Series 47, 321353
(Coagulation and Flocculation).
Snoswell, D.R.E., Duan, J., Fornasiero, D., Ralston, J., 2003. Colloid stability and the
inuence of dissolved gas. Journal of Physical Chemistry B 107 (13), 29862994.
Solomontsev, Y., White, L.R., 1999. Microscopic drop proles and the origins of line
tension. Journal of Colloid and Interface Science 218 (1), 122136.
Song, S., Lopez-Valdivieso, A., Reyes-Bahena, J.L., Lara-Valenzuela, C., 2000. Floc
otation of galena and sphalerite nes. Minerals Engineering 14 (1), 8798.
Spielman, L.A., Goren, S.L., 1970. Capture of small particles by London forces from
low-speed liquid ows. Environmental Science and Technology 4 (2), 135140.
Subrahmanyam, T.V., Forssberg, K.S.E., 1990. Fine particles processing: shear-
occulation and carrier otation a review. International Journal of Mineral
Processing 30 (34), 265286.
Sutherland, K.L., 1948. Physical chemistry of otation XI. Kinetics of the otation
process. Journal of Physical and Colloid Chemistry 52, 394425.
Sutherland, K.L., Wark, I.W., 1955. Principles of otation, Australian Institute of
Mining and Metallurgy, Melbourne, Australia.
Takahashi, M., 2005. Potential of microbubbles in aqueous solutions: electrical
properties of the gaswater interface. Journal of Physical Chemistry B 109 (46),
2185821864.
Trahar, W.J., Warren, L.J., 1976. The otability of very ne particles a review.
International Journal of Mineral Processing 3 (2), 103131.
Verwey, E.J.W., Overbeek, J.T.G., 1948. Theory of the Stability of Lyophobic Colloids.
Elsevier, New York. p. 216.
Wang, Y.Y., Betelu, S., Law, B.M., 1999. Line tension effects near rst-order wetting
transitions. Physical Review Letters 83 (18), 36773680.
Warren, L.J., 1984. Ultrane particles in otation. In: Symposia Series Australasian
Institute of Mining and Metallurgy, vol. 40 (Princ. Min. Flotation), pp. 185213.
Warren, L.J., 1992. Shear occulation. Developments in Minerals Processing 12,
309329 (Colloid Chemistry Mineral Processing).
Weisenberger, S., Schumpe, A., 1996. Estimation of gas solubilities in salt solutions
at temperatures from 273 K to 363 K. AIChE Journal 42 (1), 298300.
Widom, B., 1995. Line tension and the shape of a sessile drop. Journal of Physical
Chemistry 99 (9), 28032806.
Yang, J., Duan, J., Fornasiero, D., Ralston, J., 2003. Very small bubble formation at the
solidwater interface. Journal of Physical Chemistry B 107 (2), 61396147.
Yang, S.-M., Han, S.P., Hong, J.J., 1995. Capture of small particles on a bubble
collector by Brownian diffusion and interception. Journal of Colloid and
Interface Science 169 (1), 125134.
Ye, Y., Miller, J.D., 1988. Bubble/particle contact time in the analysis of coal otation.
Coal Preparation (London, United Kingdom) 5 (34), 147166.
Yoon, R.H., Luttrell, G.H., 1986. The effect of bubble size on ne coal otation. Coal
Preparation 2, 179192.
Yoon, R.H., Luttrell, G.H., 1989. The effect of bubble size on ne particle otation.
Mineral Processing and Extractive Metallurgy Review 5, 101122.
Yoon, R.H., Yordan, J.L., 1991. The critical rupture thickness of thin water lms on
hydrophobic surfaces. Journal of Colloid and Interface Science 146 (2), 565572.
Yoon, R.H., Mao, L., 1996. Application of extended DLVO theory, IV. Derivation of
otation rate equation from rst principles. Journal of Colloid and Interface
Science 181 (2), 613626.
Young, T., 1805. An essay on the cohesion of uids. Philosophical Transactions of the
Royal Society of London 95, 6587.
Zhou, Z.A., Xu, Z., Finch, J.A., 1994. On the role of cavitation in particle collection
during otation a critical review. Minerals Engineering 7 (9), 10731084.
Zhou, Z.A., Xu, Z., Finch, J.A., 1995. Fundamental study of cavitation in otation. In:
Proceedings of the 19th International Mineral Processing Congress, San
Francisco, vol. 3, pp. 9397.
Zhou, Z.A., Xu, Z., Finch, J.A., Hu, H., Rao, S.R., 1997. Role of hydrodynamic cavitation
in ne particle otation. International Journal of Mineral Processing 51 (14),
139149.
T. Miettinen et al. / Minerals Engineering 23 (2010) 420437 437

S-ar putea să vă placă și