You are on page 1of 12

International Journal of Mineral Processing 156 (2016) 75–86

Contents lists available at ScienceDirect

International Journal of Mineral Processing

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

A review of stochastic description of the turbulence effect on


bubble-particle interactions in flotation
Anh V. Nguyen a,⁎, Duc-Anh An-Vo b, Thanh Tran-Cong b, Geoffrey M. Evans c
a
School of Chemical Engineering, The University of Queensland, Brisbane, QLD 4072, Australia
b
Computational Engineering and Science Research Centre, The University of Southern Queensland, Toowoomba, QLD 4350, Australia
c
School of Engineering, The University of Newcastle, Callaghan, NSW 2308, Australia

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

Article history: Flotation in mechanically agitated cells has been the workhorse of the mining industry, but our quantitative
Received 28 February 2016 understanding of the effect of microturbulence generated by agitation on flotation is still very limited. This
Received in revised form 8 May 2016 paper aims to review the literature on quantifying the microturbulence effects on bubble-particle interactions
Accepted 9 May 2016
in flotation. The particular focus is on the stochastic description of bubble-particle interactions in the turbulent
Available online 11 May 2016
flow which is a random field. We briefly review the stochastic description of microturbulence and motions of
Keywords:
particles of micrometre sizes and bubbles of millimetre sizes in the isotropic turbulence of mechanical flotation
Collision cells. The key starting point is the generic equation of motion, which can be decomposed into the mean turbulent
Detachment variables and fluctuating turbulent variables. The turbulent flow of the carrying liquid is characterised using
Turbulence isotropic turbulence theory. The next focus is on reviewing bubble-particle turbulent collision and detachment
Mixing interactions. Bubble-particle turbulent collision is poorly quantified; no quantitative models of the bubble-
Minerals particle turbulent collision efficiency relevant for flotation are available. Current theories on bubble-particle
turbulent detachment face some deficiencies. In assessing the microturbulence effect on bubble-particle
detachment, the majority of studies only considers the particle acceleration in the centrifugal direction but ignore
the transverse acceleration of particles, which is due to turbulent shear flow. Critically, contact angle required in
quantifying the detachment is not constant, single-valued as considered in the theories, but can vary from
receding to advancing value during the relaxation of the triple contact line on the particle surface. The latest
experiments show that multiple-valued contact angle can significantly affect stability and detachment of floating
particles. Finally, quantifying the microturbulence effect on flotation requires further research.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction macro- and micro-processes requires the application of the statistical


turbulence theory as discussed by many flotation scientists, including
Froth flotation is an industrially important process which has been Professor Schubert et al. (Schubert and Bischofberger, 1978). The signif-
used for more than 100 years to recover valuable minerals from the icant effect of turbulence on the flotation macro-process is usually con-
Earth's crust (Lynch et al., 2010). It uses the differences in the affinity sidered via the turbulent transport phenomena, including suspension
(hydrophobicity) of solid particles to air bubbles rather than water. and transport of solid particles to avoid the sanding problem occurring
Froth flotation is usually preceded by crushing and grinding the ore to at the bottom of the cells and turbulent entrainment of solid particles
liberate the valuable mineral particles from the rock. The slurry is from the pulp phase into the froth phase (Guerra and Schubert, 1997).
mixed with flotation reagents to control the difference in hydrophobic- Turbulence can significantly affect the following flotation micro-
ity between the wanted and unwanted particles. Air is then introduced processes (Schubert and Bischofberger, 1998; Nguyen and Schulze,
into the slurry and dispersed into fine bubbles of 1–3 mm in diameter 2004):
by turbulence generated by agitation. The air bubbles collect the
hydrophobic (wanted) mineral particles and rise to the slurry surface, - turbulent dispersion of air into fine bubbles
where they form a froth layer, overflowing the lip of the cell into the - turbulent bubble-particle collision, and
concentrate launder. This very simplistic description of the widely- - turbulent bubble-particle detachment.
used flotation process in the mineral industry highlights the important
role of turbulence in flotation. Indeed, the modelling of the flotation
The turbulent transport phenomena of flotation are controlled by
⁎ Corresponding author. macro-turbulence which is defined by the size and geometry of turbu-
E-mail address: anh.nguyen@eng.uq.edu.au (A.V. Nguyen). lence-generating devices and is not considered in this review. The

http://dx.doi.org/10.1016/j.minpro.2016.05.002
0301-7516/© 2016 Elsevier B.V. All rights reserved.
76 A.V. Nguyen et al. / International Journal of Mineral Processing 156 (2016) 75–86

flotation micro-processes are significantly controlled by micro-turbu-


lence, which is universal and does not depend on the size and geometry
of turbulence-generating devices. This review focuses on the literature
which describes the effect of microturbulence on flotation.

2. Isotropic turbulence

Turbulent flow is characterised by random fluctuations in the flow


variables as illustrated for the turbulent velocity of fluid particles in
Fig. 1. Therefore, turbulent flows require statistical techniques for their
analysis, and the theory of turbulent flow is statistical in nature. Both
the mean (time-averaged) and fluctuating turbulent variables are
relevant for analysing flotation micro-processes which are controlled
by microturbulence. The flow velocity, stress and pressure (ϕ) are
decomposed into the mean (ϕ) and fluctuating (ϕ′) components by ϕ ¼
ϕ þ ϕ0 . Substituting these variables into the instantaneous continuity
and Navier-Stokes equations and applying time averaging yield the Fig. 2. The spectrum of the turbulent energy versus the reciprocal of turbulent eddy sizes.
Reynolds-averaged Navier-Stokes equations for the mass and momen- The largest eddies depend on the geometry of the turbulence-generating systems. The
tum conservations which are useful for further analysis. For instance, medium-size eddies make the main contribution to the kinetic energy. The small eddies
are isotropic, independent of the conditions of formation, as described by the universal
the statistical turbulence theory of Kolmogorov (Kolmogorov, 1941)
equilibrium range. The eddy Reynolds number of the smallest eddies at the bottom of
shows that microturbulence (i.e., turbulence of the small eddies in the the eddy spectrum is equal to 1 (Nguyen and Schulze, 2004).
universal equilibrium range as illustrated in Fig. 2) depends solely on
the turbulent kinetic energy dissipation rate, ε, and the kinematic
viscosity, ν, as follows: a vector and can be resolved into the random scalars in the appropriate
coordinates, both the longitudinal (Dll) and transverse (Dnn) structure
 1=4 functions are obtained by the projections to vector r and its perpendic-
λK ¼ ν3 =ε ð1Þ
ular direction. Dll and Dnn are linked by the following equation derived
from the mass conservation (continuity) equation:
W K ¼ ðενÞ1=4 : ð2Þ
r dDll
Dimensional analysis shows that the size, λ, and velocity, Wλ, of tur- Dnn ðr Þ ¼ Dll ðr Þ þ : ð4Þ
2 dr
bulent eddies in the equilibrium range is correlated by Wλ = (λε)1/3.
The Reynolds number of the turbulent eddies, Reλ ≡ λWλ/ν, of the
It is, therefore, sufficient to consider the structure function in the
Kolmogorov scales is equal to 1. For large eddies, Reλ is large, and dissi-
longitudinal direction, which is customarily described by its square
pation by viscous forces is unimportant. Reλ decreases with decreasing pffiffiffiffiffiffi
the eddy scale. At a critical scale, the inertial forces of turbulent eddies root as follows: ΔW ≡ Dll .
are balanced by the viscous forces, which starts affecting the motion For the smallest eddies r ~ λK, the asymptotic analysis of the
of turbulent eddies. At the critical scale, Reλ = 1 as per the Kolmogorov Kolmogorov theory gives
theory. The typical range of microturbulence is shown in Table 1.
 ε 1=2
The fluctuating turbulent velocities W 0 ¼ W−W cannot be used to ΔW ðr Þ ¼ r: ð5Þ
characterize microturbulence since they are conditioned by the large 15ν
turbulent eddies. Microturbulence can be characterised by the second
statistical moment of the velocity increment, D(r) (Kolmogorov, For large eddies, λK bb r bbΛ (macro-turbulence), we have:
1941). It is defined by the variance of the fluctuating velocities between pffiffiffi
close points at distance r as follows: ΔW ðr Þ ¼ C ðεrÞ1=3 ð6Þ

 2
Dðr Þ ≡ W 0 ðxþr; t Þ−W 0 ðx; tÞ ð3Þ where C = 2 is a numerical constant that is universal in the Kolmogorov
theory. For the whole range of microturbulence, we can expect the
where x is the position vector and r = | r |. D(r) is also known as the following expression:
structure function of the random scalar field W(r). Because velocity is
ΔW ¼ W K f ðr=λK Þ ð7Þ

where f is an interpolating function, connecting the two asymptotic


expressions described by Eqs. (5) and (6). The asymptotic analysis
(Batchelor, 1982) gives the following interpolating function valid in

Table 1
Scales of the smallest laminar and turbulent eddies in water as a function of the energy
dissipation rate ε (ν = 10−6 m2/s).

Dissipation rate ε Scale of the smallest


[W/kg]
Laminar eddies λ = (4–6) Turbulent eddies λ = (10–15)
λK [μm] λK [μm]

1 130–190 320–470
10 70–110 180–270
Fig. 1. Schematic of the actual, time-averaged and fluctuating velocities of turbulent flow
100 40–60 100–175
as a function of time (Nguyen and Schulze, 2004).
A.V. Nguyen et al. / International Journal of Mineral Processing 156 (2016) 75–86 77

Fig. 3. Root-mean-square of the fluctuating velocities of microturbulence in the longitudinal direction between two points separated by a distance r. Eq. (8) is described by the thick line.
The thin lines describe the appropriate subranges.

the whole universal equilibrium range (Fig. 3): within the range of 3 to 4 times of the Kolmogorov microscale. In flota-
pffiffiffiffiffiffi tion, they are within the range of about 50 to 100 μm (Table 1). Fine
x= 15 solid particles of the sizes within the range up to 10λK are influenced
f ðxÞ ¼ h i1=3 : ð8Þ
mainly by laminar eddies in the dissipative subrange. Air bubbles of
1 þ ð15C Þ−3=2 x2
sizes larger than 10λK are influenced by the inertial eddies.
The pressure and acceleration of turbulent flows are also critical
Microturbulence is often divided into the dissipative and inertial to bubble-particle interactions. Similar to the spatial correlations
subranges (Fig. 3). The dissipative subrange is used for r/λK ≤ 5 to 10. between turbulent velocities, the turbulent pressure fluctuation can be
The root-mean-square of fluctuating velocities in this subrange is de- characterised by the spatial structure function which is defined by
scribed by Eq. (5). Eddies in this subrange are laminar. The maximal
Dpp ðrÞ ≡ ½P 0 ðxþr; tÞ−P 0 ðx; tÞ . The root mean square of the pressures is
2
laminar shear stress in this subrange is given by pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
often used, giving ΔPðrÞ ¼ Dpp ðrÞ . It can be described in terms of
pffiffiffiffiffiffi
τ max ¼ 0:26δ εν: ð9Þ ΔW expressed by Eq. (7) or its universal function f(x) as follows
(Batchelor, 1982):
At 0.06Λ ≥ r/λK ≥ 15 to 20 the inertial transfer of energy from large vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u ! !
eddies is the dominant factor, so this is called the inertial subrange. u
2 Z∞ 2 2 ZK
r=λ
2 2
u r 1 df df
The root-mean-square of the difference in the turbulent velocities ΔP ðr Þ ¼ δW K 2 u
t λ dx þ x dx: ð11Þ
between sufficiently near points separated by a distance r in this K x dx dx
pffiffiffi r=λK 0
subrange is approximated by Eq. (6) with C ¼ 2. Eddies of these scales
are turbulent in themselves. The fluctuation of the normal stresses in
Upon inserting Eq. (8) for f, Eq. (11) can be numerically integrat-
this subrange is dominant and strongly controls a number of physical
ed. The result of the integration is shown in Fig. 4. For large eddies,
processes occurring in the liquid, including the air dispersion, the bub- pffiffiffi
f ¼ C x1=3 as per Eqs. (7) or (8), and Eq. (11) gives
ble-particle collision (Not all the bubble-particle collision models are
applicable for particles of size of random eddies, as the comparative
ΔP ðr Þ ¼ δCðεrÞ2=3 ¼ δðΔW Þ2 : ð12Þ
contributions from the inertial, gravitational and interceptional effects
are dependent on the particle size) and the stability of bubble-particle
aggregates in flotation. The root mean square of turbulent pressures Eq. (12) is the well-known prediction for spatial pressure change in
(or “dynamic thrust“) in the inertial subrange is determined by Eq. (12). the inertial subrange (λK ≪ r ≪Λ), which is widely used in predicting
As some microturbulence processes occur in the range r/λK = 5 to the breakage of air bubbles.
30, Schubert et al. (Schubert et al., 1990) introduced a transitive The Kolmogorov method of microturbulence analysis can be used
subrange between the dissipative and inertial subranges. The square to analyse the fluctuation of turbulent acceleration. The period of oscilla-
root of the velocity variance in this transitive subrange is described as tion of an eddy λ is equal to Tλ ∼λ/Wλ =ε2/3λ−1/3. The eddy acceleration,
follows: aλ, follows the proportionality: aλ ≡dWλ/dTλ ∼Wλ/Tλ ∼ε2/3λ−1/3. The eddy
acceleration has two components, as described by the Navier-Stokes
ΔW ¼ 0:45ε5=12 ν −1=4 r 2=3 : ð10Þ equations, the potential acceleration component produced by pressure
fluctuation and the solenoidal component produced by viscous forces.
This transitive subrange is valid for 5λK ≤ r ≤ 29λK. Therefore, the spatial correlation of the acceleration field is a tensor
A common feature of turbulent flows in flotation is the small size of which can be described as follows: aðx þ r; tÞaðx; tÞ ¼ ½Bll ðrÞ−Bnn ðrÞr i r j =
particles and bubbles in comparison with the dimension of the turbu- r 2 þ Bnn ðrÞδij , where ri and rjare the moduli of position vectors, δij is the
lence macroscale. The small eddies are determined by the dissipation Kronecker delta, and Bll(r) and Bnn(r) are the longitudinal and transverse
of local turbulent energy and the liquid kinematic viscosity, and are correlation functions of the acceleration field. These correlation functions
78 A.V. Nguyen et al. / International Journal of Mineral Processing 156 (2016) 75–86

know very little about the exact motions of air bubbles and solid
particles in flotation, especially in the turbulent flows of flotation.
Consequently, a number of approximate approaches have been applied
and useful information has been obtained. A popular approach is to
adopt the extensions of the Basset-Boussinesq-Oseen (BBO) equation
for the particulate motions in non-stationary fluid (Nguyen and
Schulze, 2004). The BBO equation includes steady and unsteady drag,
gravitational and inertial forces. For fine particles typically encountered
in flotation the extended BBO equation can be described as follows
(Nguyen and Schulze, 2004):

dV δ DW δ 1 dV dW
mp ¼ mp −mp − −6πμRp ðV−WÞ
dt ρp Dt ρ 2 dt dt
! p
δ
þ mp 1− g ð16Þ
ρp

where D(...)/Dt = ∂(...)/∂ t + W ⋅ grad(...) and d(...)/dt = ∂(...)/


∂ t + V ⋅ grad(...) describe the derivatives following the liquid element
and the moving solid particle, respectively. Here, the history Basset
Fig. 4. Root-mean-square of the fluctuating pressures of microturbulence between two term of the dynamic resistance is neglected. Eq. (16) is applied to very
points separated by a distance r. small particles that ideally satisfy Stokes Law. Empirically, it can be
used with particles with a diameter up to 60 μm. For coarse particles,
corrections are needed but the most important extension is the correc-
can be derived from the Navier-Stokes equations (Batchelor, 1982) tion to the Stokes resistance by multiplying the third term on the right-
and yield, for the inertial subrange, where viscous contributions are hand side of Eq. (16) by a factor of fd = 1 + 0.169Re2/3 which is valid for
negligible: 0 ≤ Re b 700 (Nguyen and Schulze, 2004). Alternatively, the factor of the
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 2
Oseen correction can be used but is only valid for very small Reynolds
1 d Dpp ðr Þ 8 ε4=3 number. These two correction factors can permit analytical solutions
Bll ðr Þ ¼ al ðx þ r; t Þal ðx; t Þ ¼ 2 ¼ 2=3 ð13Þ
2δ dr2 9r of the particle motion equation. The Reynolds number is calculated
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi using the local slip velocity.
1 dDpp ðr Þ 8 ε4=3 The description of the motion of air bubbles in flotation is more
Bnn ðr Þ ¼ an ðx þ r; t Þan ðx; t Þ ¼ 2 ¼ 2=3 ð14Þ
2δ r dr 3r complicated than that of solid particles. The bubble motion is influenced
by the bubble size and shape, physical properties of the medium, and
where Dpp(r)=[ΔP(r)]2 =δ2C2(εr)4/3 =4δ2(εr)4/3. The magnitude of the physicochemical properties of the gas-liquid interface (Nguyen and
spatial correlation of eddy acceleration, often referred to as the mean Schulze, 2004). The shape that the bubbles assume during their motion
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
“machine acceleration”, can be determined by: bm ¼ Bll ðrÞ þ Bnn ðrÞ , is, in turn, a complex function of the hydrodynamic, viscous and interfa-
which gives cial forces exerted on the bubbles. The liquid drag force on small bub-
bles may be equal to that on solid particles. The bubble surface may
ε2=3 become mobile, and in contrast to that for solid particles, there are
bm ðrÞ ¼ 1:9 : ð15Þ
r 1=3 non-vanishing tangential velocities at the gas-liquid interface. The gas
circulation inside a bubble can also reduce the drag on it. Importantly,
The machine acceleration was used in quantifying bubble-particle flotation reagents and surface contaminants can have a significant effect
detachment (Schulze, 1982). on the bubble hydrodynamics. They can adsorb onto the bubble surface,
One of the most important parameters determining the turbulent retard the bubble mobility and restore the shape of small bubbles to a
detachment is the rate of energy dissipation. Unfortunately, a numerical sphere or spheroid. The drag force on air bubbles rising through a
value of this parameter is not always available, and a mean dissipation flotation pulp usually follows the drag force on bubbles in grossly
rate is often used. The mean rate is defined as the input power per contaminated water. Therefore, the following equation can be used to
unit mass of liquid. It is rather a rough estimation for flotation because describe the motion of air bubbles in the turbulent flow of flotation:
of the spatial distribution of the actual dissipation in flotation cells. In

the impeller-stator region the gradient of the dissipation rate of turbu- dU δ DW δ 1 dU dW
lent energy is the highest. The maximum value of the local dissipation in mb ¼ mb −mb − −6πμRb ðU−WÞf d
dt ρb Dt
ρb 2 dt dt
the impeller region of industrial flotation cells can be higher than the δ
þ mb 1− g ð17Þ
mean value by a factor up to 30. In this region the total power input dis- ρb
sipates, and the dispersion of air into bubbles and the bubble-particle
interaction actively occur. The mean turbulent energy dissipation rate where d(...)/dt=∂(...)/∂t+U⋅grad(...) describe the derivatives following
obtained by averaging over the total cell volume can be incorrectly the moving bubble. Eq. (17) is similar to Eq. (16), except for the correction
used for the modelling of bubble-particle interactions and other factor, fd = 1 + 0.169Re2/3, for the drag force which must be included
micro-processes in flotation. A good estimation should be based on because of the millimetre size of air bubbles used in flotation. Here, the
the input power and the volume of the impeller region, where the local bubble Reynolds number is defined as Re=2Rb |U−W|/v. The histo-
turbulent bubble-particle interactions take place. ry Basset term of the dynamic resistance is also neglected.
Both Eqs. (16) and (17) show that the motions of solid particles
3. Motion of particles and bubbles in turbulent flows and and air bubbles are controlled by the gravitational, inertial and liquid
their interactions flow forces. These forces can have strong effects on bubble-particle
interactions, and are usually included in the predictions of bubble-
Bubble-particle interactions in flotation can be quantified if their particle interactions. Their effects on bubble-particle interactions in
motions in the flotation cell are known. Regrettably, we almost always flotation are usually coupled. Therefore, the final effect of these forces
A.V. Nguyen et al. / International Journal of Mineral Processing 156 (2016) 75–86 79

on bubble-particle collision, attachment and detachment interactions is 4. Effect of turbulence on bubble-particle collision interaction
difficult to predict by linear combinations of the individual effects. The
exact solutions of the motion equations are not always attainable. The The theories of bubble-particle collision (encounter) interaction can
mobility of the bubble surface can be introduced into quantifying the be deterministic or stochastic. These theories aim to predict the collision
bubble-particle interactions through the boundary conditions at the rate and efficiency. The available theories are based on the assumption
air-water interface applied to the water velocity. As can be seen from that motions of bubbles and particles in flotation are deterministic.
Eqs. (16) and (17), both the particle and bubble velocities required for Consequently, these theories are evidently not able to describe the
quantifying bubble-particle interactions can be calculated if the water collision interactions affected by fluctuating components of turbulence
velocity is known. Therefore, the modelling of the bubble-particle inter- which are not deterministic. The available deterministic theories,
actions is usually preceded by the modelling of the water flow field and however, can be applied to describe the collision interactions controlled
then the modelling of the particle and bubble velocities using the BBO by the mean components of turbulent flow which are determined by
equations. large (energy-containing) turbulent eddies and turbulence-generating
The motion equations for both the solid particles and air bubbles in devices.
turbulent flows, described by Eqs. (16) and (17), respectively, are non- The bubble-particle collision rate, Nc, is defined as the number of
linear and therefore difficult to be solved analytically. Some approxima- particles colliding with a bubble per unit time, and can be described as
tions have been obtained. For example, in the case of fine particles follows:
which are entrained in the turbulent water flow, after applying the
Reynolds decomposition, Eq. (16) can be transformed to the following Nc ¼ ∮ J  dSb ¼ ∮ np ΔV dSb ð21Þ
equation for the fluctuating variables:
where J = npΔVis the flux of the free (unattached) particles moving rel-
0 0 atively to the bubble with a relative velocity ΔV, np the particle number
dV ðt Þ 1 0 3δ dW ðt Þ 1 0
þ V ðt Þ ¼ þ W ðt Þ ð18Þ density (concentration), and the surface element vector dSb is the vector
dt T 2ρp þ δ dt T
of the bubble surface normal—defined as one particle radius away from,
and pointing towards the bubble surface. Eq. (21) can be applied to
where the relaxation time, T, of the particle with a “joined mass” is modelling both the deterministic and stochastic collision interactions
described by in flotation. Utilising the literature involving turbulent gas-particle
flows and turbulent coagulation (Williams and Crane, 1983; Kruis and

Kusters, 1997), the collision rate is defined by the collision frequency
2 1
T¼ ρp þ δ Rp 2 : ð19Þ function (or kernel), β, yielding
9μ 2
Nc ¼ np β ð22Þ
It is noted that V−W ¼ 2ðρp −δÞgR2p =ð9μÞ is the Stokes velocity for
where β is the number of collisions per unit time and volume.
particle settling and is utilised in establishing Eq. (18). Employing the
Collision efficiency, Ec, is defined as the ratio of the real, Ncr, to ideal,
initial condition: V′(0) = W′(0) = 0, Eq. (18) can be solved to obtain
Nci, collision rates calculated based on Eq. (21) for the particle motions
the following prediction for the particle fluctuating velocity:
towards the bubble: Ec = Ncr/Nci. In calculating the ideal collision rate,
Z Nci, the particle motion is considered not to be influenced by the
3δ 2ρp −2δ 1 t
τ
V0 ðt Þ ¼ W0 ðt Þ þ e−T W0 ðt−τÞdτ: ð20Þ presence of the bubble, yielding (Nguyen and Schulze, 2004)
2ρp þ δ 2ρp þ δ T 0
 2  
Nci ¼ πnp Rp þ Rb V p þ U b ð23Þ
Eq. (20) provides a special relationship between the fluctuating
velocities of the water and fine solid particles which are entrained in where Vp and Ub are the terminal velocities of particle settling and bub-
the water turbulent flow. They are the random variables but can be ble rising, respectively. Eq. (23) links the bubble-particle interactions
used for demonstrating the effect of water turbulence on bubble-particle with flotation kinetics based on the first principle (Nguyen and
collision interaction. Schulze, 2004). In flotation, solid particles are small, and their motion
Particle-bubble interaction underpins the flotation process, and is around the bubbles is solenoidal. The particle concentration in Eq.
normally divided into collision, attachment and detachment (Jameson (21) can be taken outside the integral and we obtain the following
et al., 1977; King, 2001; Nguyen and Schulze, 2004). Collision is the generic equation for collision efficiency:
approach of a particle to encounter a bubble and is governed by the
∮ ΔV dSb β
mechanics of the bubbles and particles in the turbulent flow field. The Ec ¼  2  ¼  2   ð24Þ
limit of the collision process is determined by the zonal boundary π Rp þ Rb U p þ U b π Rp þ Rb U p þ U b
between the long-range hydrodynamic and interfacial force interac-
tions (Nguyen and Schulze, 2004). The inter-surface separation distance where β ¼ ∮ ΔV dSb is obtained by comparing Eqs. (21) and (22) under
at the zonal boundary between a bubble and a particle is of the sub- the solenoidal condition. The calculation of the integral in Eq. (24)
micrometre order. Once the particle approaches the bubble at a shorter is central to developing the theories of bubble-particle collision
separation distance, the molecular forces are significant, and the interaction.
attachment process starts. The detachment process is governed by the In flotation cells the bubbles rise to the surface whilst the particles
capillary force, the particle weight and the detaching forces of the turbu- settle to the bottom unless they are captured by the bubbles. The pro-
lent flow. The governing mechanisms of long-range hydrodynamics, cess can generally be considered as a counter-current operation with
surface chemistry, and capillarity are largely independent, and each of the collision interaction occurring at the upper part of the bubble sur-
them has a significant influence on only one of the three elements of face where the particles approach the bubble. Therefore, the integration
the bubble-particle interaction. Consequently, collision, attachment in Eq. (24) is limited by the bubble equator. The collision interaction
and detachment can be mathematically treated independently, which cannot occur at the bubble equator, but is limited by a polar angle, φc,
significantly simplifies the task of modelling each of the interactions. (measured from the upper pole of bubble surface) smaller than 90
φ
The following sections focus on the effect of turbulence on the collision degrees, leading to Ec ¼ 2∫ 0 c ð−er  ΔVÞ sinφdφ=ðV p þ U b Þ, where er is
and detachment interactions. the unit vector of the bubble surface normal. The deviation of φc from
80 A.V. Nguyen et al. / International Journal of Mineral Processing 156 (2016) 75–86

90 degrees is due to a number of reasons, including the fore-and-aft velocity, which is derived from the probability distribution function
asymmetry of water flow around air bubbles (Dobby and Finch, 1986; (PDF), pðΔV r 0 Þ, and is given by:
Nguyen and Schulze, 2004), centrifugal forces (Dai et al., 1998) and
  Z þ∞  
inertial forces (Ralston et al., 1999; Nguyen and Schulze, 2004). The ΔV r 0 ¼ ΔV r 0 p ΔV r 0 dΔV 0 ð28Þ
r
fore-and-aft asymmetry is due to the millimetre size of the air bubbles −∞
used in flotation, which have a Reynolds number, Re, usually in the
range 1–500. For this intermediate range of Reynolds numbers the non- where pðΔV r 0 ÞdΔV r 0 is the probability of ΔV r 0 taking a value between Δ
linear (viscous) term in the Navier-Stokes equations is important and V r 0 and ΔV r 0 þ dΔV r 0 . The PDF is usually assumed to follow a Gaussian
can cause the water streamlines to become more compressed towards probability distribution
the upper bubble surface than the lower surface. The fore-and-aft asym- !
  1 ΔV r 0 2
metry is more significant for bubbles with an immobile surface than p ΔV r 0 ¼ pffiffiffiffiffiffi exp − ð29Þ
bubbles with a mobile surface. Since the Stokes and potential flows σ 2π 2σ 2
are fore-and-aft symmetric their linear combination does not predict
and represent the fore-and-aft asymmetric water flow around air where σ 2 ¼ hΔV r 0 2 i is the variance (second moment) of ΔV r 0 . Inserting
bubbles in flotation realistically and, therefore, the deterministic Eq. (29) into Eq. (28) and integrating gives
theories (Yoon and Luttrell, 1989; King, 2001) developed based on the rffiffiffi
simple linear combination of the Stokes and potential flows are evident-   2
ΔV r 0 ¼ σ ð30Þ
ly of artefact. These complicated details have led to many simplified π
deterministic theories. The model developments are covered by a num-
ber of reviews (Dai et al., 2000; Phan et al., 2003; Kostoglou et al., 2006; The collision frequency function and efficiency due to turbulence
Nguyen, 2011) and are not repeated here. Table 2 shows a summary of fluctuation are directly proportional to the second moment of the radial
the available deterministic models which are helpful for the discussion fluctuating relative velocity of particles. Inserting Eq. (30) into Eq. (27),
in this paper. the fluctuating collision frequency is given by:
Regarding bubble-particle collision interactions in turbulent flows, a pffiffiffiffiffiffi 2
stochastic modelling approach must be adopted. The flotation literature β0 ¼ 8π Rp þ Rb σ: ð31Þ
on this topic is very limited when compared with the extensive litera-
ture dealing with turbulent gas-particle flows and turbulent coagulation The calculation of the variance of the fluctuating relative velocity of
of fine particles. Some key features found in the flotation literature are colliding particles is widely reported in the literature and recently
described below: reviewed (Meyer and Deglon, 2011). The available literature is rich
For turbulent bubble-particle collision interaction, the bubble-particle but often confusing; especially since the majority of the predictions
relative velocity can be decomposed into the time-averaged, ΔV for the fluctuating collision function for gas-particle flows and particle
(deterministic), and fluctuating (stochastic), ΔV′, components of the coagulation in liquids are not physically consistent with the bubble-
relative velocity, yielding particle interactions in flotation. The following comments are offered:

ΔVðt Þ ¼ ΔV þ ΔV0 ðt Þ ð25Þ 1) Both spherical and cylindrical formulations have been used to calcu-
late the turbulent collision frequency function, but models based on
Substituting Eq. (25) into Eq. (24) we obtain the cylindrical formulation are erroneous. The spherical formulation
gives rise to Eq. (27) and should be used while the cylindrical formu-
  dSb
∮ ΔV
0
∮ ΔV ðt Þ  dSb lation uses a cylindrical volume passing through a reference particle
Ec ¼  2  þ  2  : ð26Þ per unit time and yields β′ = π(Rp + Rb)2〈| ΔV′|〉. The differences
π Rp þ Rb V p þ U b π Rp þ Rb V p þ U b
between the two formulations include not only the numerical factor
of 2 but also the average velocities. In the most cited paper by
The collision efficiency now has two terms. Firstly, the first term on Saffman and Turner (Saffman and Turner, 1956), both formulations
the right-hand side of Eq. (26) represents the efficiency due to the time- were used, leading to different results which could not be reconciled.
averaged (mean) interactions (If the time-averaged flow variables are These two formulations were shown to produce the same results
different from zero, the results presented in Table 1 and discussed in only under special cases (Wang et al., 1998).
the preceding section can be applied). Secondly, the second term is 2) There is a similarity of mechanisms governing the deterministic and
due to the fluctuating relative motion between the bubble and particle stochastic collisions. The mechanisms of stochastic collisions include
(Likewise, the collision frequency β can also have two similar terms). diffusion, shear, accelerative, differential sedimentation and prefer-
Since ΔV′(t) is a random field, the fluctuating collision interaction ential concentration. These mechanisms follow the different forces
must be determined by a stochastic approach. Although the fluctuating governing the particle motions as described by Eq. (16). Brownian
velocities are random, the mass balance over the bubble surface diffusion is known to govern flotation of nanoparticles (Nguyen et
requires that the overall particle influx and outflux over the bubble al., 2006), and is not important for flotation processes used in the
surface be equal. The collision frequency due to fluctuating motions, mineral industry. The shear mechanism governs the particle collision
β′, is therefore half the surface area multiplied by the average magni- of particles which follow water streamlines and collide due to differ-
tude of the relative radial velocity, i.e., ent positions with the shear flow field. The shear mechanism is similar
 2   to the bubble-particle collision by interception in the deterministic
β0 ¼ ∮ ΔV0 ðt Þ dSb ¼ 2π Rp þ Rb ΔV r 0 ð27Þ models in Table 1. The differential sedimentation mechanism is anal-
ogous to the gravitational collision while the accelerative collision is
where ΔV r 0 ¼ R  ΔV0 =R is the radial component of the fluctuating rela- similar to the inertial models of the deterministic collision.
tive velocity, R is the vector of particle centres, and R = |R|. The angle 3) Unlike the deterministic bubble-particle collisions, the stochastic
brackets in Eq. (27) describe the statistical average over the colliding collisions are significantly controlled by correlation of particle
sphere which is defined, relative to the bubble, as a sphere of radius motions in the turbulent flow field (Fig. 5). For small particles
R= Rp + Rb. The calculation of the averaged velocity in Eq. (27) is central their fluctuating velocities are correlated or partially correlated
to many theories on turbulent particulate interactions. Principally, the with the velocity of carrier fluid. These small particles can collide
averaged velocity can be determined from the first moment of the by either shear flow, difference in accelerations, and gravity. The
A.V. Nguyen et al. / International Journal of Mineral Processing 156 (2016) 75–86 81

Table 2
Deterministic models for bubble-particle collision efficiency as organised according to year of their publication and their analytical-to-empirical nature (Nguyen, 2011).

Author(s) Collision efficiency, Ec Notes and symbol definitions

Sutherland (1948) Ec = 3R Interceptional effect; potential flow. R is the interception number.


where R = Rp/Rb
Gaudin (1957) Ec ¼ 32 R2 Interceptional effect; Stokes flow.
Flint and Howarth (1971) Ec ¼ V sVþU
s Gravitational effect.
Reay and Ratcliff (1973) Ec ¼ V sVþU
s
ð1 þ RÞ2 ; sin2 φc Gravitational effect.
Anfruns and Kitchener (1977) Ec ¼ V sVþU
s
ð1 þ RÞ2 þ V sUþU 2ψc Combined gravitational and interceptional effects. ψc is the stream function
ψ calculated at r = 1 + R and φ = π/2.
Weber and Paddock (1983) Ec ¼ VUs ð1 þ RÞ2 ; sin2 φc þ Ec;i Combined gravitational and interceptional effects; intermediate bubble
2
h i Reynolds number.
3Re=16
Ec;i ¼ 3R2 1 þ 1þ0:249Re 0:56 … (*)
h i Eq. (*) for bubbles with a fully immobile surface.
Ec;i ¼ R 1 þ 2
… (**) Eq. (**) for bubbles with a fully mobile surface.
1þð37=ReÞ0:85
h i
Yoon and Luttrell (1989) 0:72
Ec;i ¼ R2 32 þ 4Re15 Interceptional effect; intermediate bubble Reynolds number; fully
immobile bubble surface.
Dobby and Finch (1987) 1) Low particle inertia, St b 0.1 Combined inertial, gravitational and interceptional effects; intermediate
Ec,0 = Ec, g + Ec,i bubble Reynolds number; immobile bubble surface; analytical results for
Ec,i is given by Weber and Paddock for immobile surface. low Stokes number, St, and numerical results for intermediate St.
Ec,g is given by Reay and Ratcliff. ρ 2
St ¼ Re
9 δR
(
78:1−7:37 ; logRe if 400NReN20 Range of application of the numerical results for intermediate inertia:
φc ¼ 85:5−12:49 ; logRe if 20NReN1 300 N Re N 20; 0.8 N St; and 0.25 N Vs/U.
85:0−2:5 ; logRe if 1NReN0:1
2) Intermediate particle inertia, St b 0.1
Ec = (Ec,0)[1.627Re0.06St0.54(Vs/U)−0.16]
Ec,0 is the low-St efficiency given in 1).
h i
Schulze (1989) E
Ec ¼ Ec;i þ Ec;g þ Ec;in 1− c;i 2 Combined gravitational, inertial and interceptional effects; intermediate
ð1þRÞ
bubble Reynolds number; immobile bubble surface; approximate results for
Ec;i ¼ V sUþU 2ψc both low and intermediate St. Model parameters (a, b and φc) are functions
Ec;g ¼ VUs ð1 þ RÞ2 ; sin2 φc of Re. φc is given by Dobby and Finch. The surface vorticity is given as
9Re=16
b 3þ
Ec;in ¼ V sUþU ð1 þ RÞ2 ðStþa
St
Þ ξ¼ 1þ0:249Re0:56
2 ;sinφc
For R ≤ 1/ξ:
2
h i
3Re=16
ψc ¼ 4R2 1 þ 1þ0:249Re 0:56

Dukhin (1983), Dai et al. (1998) Ec ¼ 3R ; sin φc ; expf− ; cosφc ½3K ‴ ð ; ln R þ 1:8Þ
2 Potential flow and bubbles with a mobile surface. Significant effect of
þ ð8−12 ; cosφc þ 4 ; cos3 φc Þ=ð3 ; sin4 φc Þg centrifugal force on collision. Valid for ultrafine particles. Expression for the
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi collision efficiency is called the Generalized Sutherland Equation (GSE).
φc ¼ acosð 1 þ β 2 −βÞ
2Rf ‴ 2ðρ−δÞURp 2
β ¼ 3K ‴ and K ¼ 9μRb
Nguyen and Schulze (2004) - Gravitational collision Individual and combined inertial, gravitational and interceptional effects;
Ec;g ¼ V sVþU
s intermediate bubble Reynolds number; immobile and mobile bubble
- Interceptional collision surface; gas hold up; approximate results for both low and intermediate
n pffiffiffiffiffiffiffiffiffiffiffiffi
2 2
ffi o Stokes number, St.
2 X þ3Y −X
Ec;i ¼ f ðRÞ V sVþU
s
½X þ Y ; cosφc;i  ; sin φc;iφc;i ¼ acos 3Y
- Inertial collision Model parameters (a, b, c, X and Y) are functions of Re and gas volume
Ec;in ¼ ðStÞ
a
−b fraction (as well as the bubble surface mobility) – see Nguyen and Schulze
ðStÞa þc
(2004).
- Simultaneous gravitational and interceptional collision
f ðRÞV s
Ec;gi ¼ V s þU ½X þ C þ Y ; cosφc;i  ; sin2 φc;gi For bubbles with an immobile surface, f(R) = R2.
n pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2 2
o
ðXþCÞ þ3Y −ðXþCÞ
φc;gi ¼ acos 3Y
For bubbles with a mobile surface, f(R) = R − R2.
- Overall collision efficiency
Ec = 1 − (1 − Ec, g)(1 − Ec,i)(1 − Ec,in)(1 − Ec, t)
C ¼ VUs 1
f ðRÞ
n
Reay and Ratcliff (1975), Jiang Ec = m(R) Empirical or computational results. Model parameters m and n are only
and Holtham (1986) available for limited conditions but n N 1 for all cases.

Saffman and Turner model (Saffman and Turner, 1956) is an


excellent description applicable to turbulent collision with a strong
correlation. For large particles, their fluctuating velocities are uncor-
related with the velocity of carrier fluid. These large particles are
thrown randomly from eddy to eddy and collide. The model of
Abrahamson (Abrahamson, 1975) is a good example of the collision
of particles with uncorrelated velocities.

Saffman and Turner (1956) used a simplified version of Eq. (16)


without the added-mass force for gas-particle flows and developed a Fig. 5. Mechanisms of a particle collision in the turbulent flow field. Small particles (red
and blue circles) have velocities correlated with water flow and collide by shear and
model for the relative velocity variance σ which has three terms due accelerative correlated mechanisms. Large particles (black circles) have velocities
to shear, acceleration and gravity. Their model was corrected using the uncorrelated with carrier fluid, can be transferred randomly from eddy to eddy and
spherical formulation (Wang et al., 1998), leading to the following collide.
82 A.V. Nguyen et al. / International Journal of Mineral Processing 156 (2016) 75–86

prediction: particles are given as follows (Nguyen and Schulze, 2004):


rffiffiffiffiffiffiffiffiffiffiffiffiffiffi 7=9

!2 *
+ !2  0 2 ε4=9 Rb δ−ρb 2=3
1 2ε δ DW 2 π δ U b ¼ 0:83 ð35Þ
σ2 ¼ R þ 1− ðτ 1 −τ 2 Þ2 þ 1− ðτ1 −τ 2 Þ2 g2 v1=3 δ
15
|fflfflfflffl{zfflfflfflfflv} ρp Dt 8 ρp
Shear
|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl} |fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl} rffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Acceleration Gravity  0 2 2R3p ε ρp −δ
!2 *
+ Vp ¼ : ð36Þ
δ DW 2 R2 135 ν2 δ
þ2 1− τ1 τ2
ρp Dt λD 2 ð32Þ
|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl} When body forces, such as the gravitational forces which drive the
Coupling particle settling and bubble rising in flotation are considered, the
Abrahamson solution can be converted to the collision frequency
where R is the sum of the two particle radii and λD is the Taylor micro- function as follows (Nguyen and Schulze, 2004):
scale of fluid acceleration. The coupling term is absent in the original 2rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi3
model and accounts for the combined effect of spatial variation of 02 02
  6 V p þ U b7
fluid acceleration and particle inertia. Since air bubbles rise against β0 ¼ R V p þ U b
2
χ6
4 Vs þ U 5
7 ð37Þ
gravity, and solid particles sink in the direction of gravity, the gravity b

term applied to the bubble-particle collision should be additive (i.e.,


τ1 + τ2) because the gravitational forces in Eqs. (16) and (17) are
where function χ is defined by χðxÞ ¼ 23=2 π1=2 x expf− 2x12 g þ πð1 þ
opposite in sign. Moreover, Eq. (32) is strictly valid for fine particles of
a size of the smallest eddies and, therefore, has to be modified for flota- x2 Þ erffp1ffiffi2xg. Since χ(0) = 1 as in the case of no turbulence (the bar
tion to account for the size of air bubbles. A number of modifications and variables are zero), the collision is expected under a constant force
extensions are available as summarised by Meyer and Deglon (Meyer field of gravity. If gravity is ignored (Vs + Ub → 0), Eq. (37) reduces to
and Deglon, 2011). For flotation, the shear term should also be extended Eq. (33).
to the inertial subrange using Eq. (7). Also, Eq. (32) has been approxi- Eq. (34) was confirmed by experiments for 80 μm solid particles but
mately developed by neglecting the particle acceleration in the motion underestimates the fluctuating velocities of 1.2 mm air bubbles (Brady
equation. This approximation can be improved or replaced by the et al., 2006). Finally, it is noted that the proposed models for bubble-
exact solution of the particle motion equation which is described by particle turbulent collision need further experimental validation.
Eq. (20); where the first term on the right-hand side can give rise to a
similar shear term in Eq. (32) and the integral term can be stochastically 5. Effect of turbulence on detachment
treated, leading to the other accelerative terms which can contain the

following (or similar) integral: ∫ 0 e−τ=T DL ðτÞdτ , where DL(τ) is the Quantifying the stability of bubble-particle aggregates in flotation
Lagrangian structure function of fluid flow. The exact approach was aims to determine if the adhesion on the attached particle is strong for
briefly described by Panchev (Panchev, 1971), but has since been supporting the particle to avoid detaching from the bubble surface by
comprehensively developed for particle turbulent collision by Zaichik the disruptive forces in flotation cells. The force analysis on the attached
et al. in their book (Zaichik et al., 2008). particle concludes that the Archimedes Principle still holds for equilibri-
Large particles exhibit significant inertia when subjected to turbu- um of the particle attachment to the air-water interface: The weight of
lent fluid flow fluctuation and have velocities uncorrelated with the liquid displaced by air (of volume enclosed by the deformed air-water
fluid. Abrahamson (Abrahamson, 1975) assumed the Gaussian proba- interface, air-solid interface, and the initial plane) is balanced by the
bility distribution function for the particle fluctuating velocities in 3D particle weight, i.e.,
space and calculated the collision rate which can be converted to the
collision frequency function as follows: Fb ¼ Fg ð38Þ

rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi where buoyancy consists of the liquid weights of volumes above the
0 3=2 1=2 2  0 2  0 2 deformed air-water interface and the cylinder above the three-phase
β ¼2 π R V1 þ V2 ð33Þ
contact area less the spherical cap. These three volumes are highlighted
in green, purple and dark blue in Fig. 6. The liquid weight of the first
where R is the sum of the two particle radii. The particle velocity volume is equal to the vertical component of capillary force (Nguyen
variances in Eq. (33) are obtained using the solution of Eq. (16) and
an exponential form for the Lagrangian velocity correlation. Alternative-
ly, the experimental correlations obtained by Liepe and Möckel (Liepe
and Moeckel, 1976) can be used. These correlations are as follows:

rffiffiffiffiffiffiffiffiffiffiffiffiffi 7=9

 0 2 ε4=9 Ri ρp −δ 2=3
V i ¼ 0:57 for i ¼ 1 and 2: ð34Þ
v 1=3 δ

Eqs. (33) and (34) are commonly used to model flotation systems
(Schubert, 1999; Pyke et al., 2003; Kostoglou et al., 2006), where Eq.
(34) was established based on the balance between the inertial
subrange acceleration (Table 2) and Allen's fluid drag on particles.
This balance is appropriate for particles and fine air bubbles in flotation
(The exact constant using Allen's drag is equal to 0.83 (Nguyen and
Schulze, 2004) which is larger by a factor of 1.5 than the constant of
the empirical correlation of 0.57). The appropriate balance for fine
particles is between dissipative subrange acceleration and Stokes' Fig. 6. Force balance on a spherical particle attached to the air-water surface shows the
drag. Therefore, the corresponding equations for fine bubbles and fine validity of the Archimedes Principle: Fb =Fg.
A.V. Nguyen et al. / International Journal of Mineral Processing 156 (2016) 75–86 83

and Schulze, 2004). It gives: bubble diameter as used by Schulze (1982), the bubble radius is used
in Eq. (42) as recommended by Goel and Jameson (2012). It was argued
F b ¼ 2πRp σ sinα sinβ that the radius of rotation (of the eddy) would seem more reasonable to
 
2 2−3 cosα þ cos3 α πRp 3 be equal to the radius of the bubble than the bubble diameter.
þ πðRP sinα Þ Hδg− δg: ð39Þ
3 The machine acceleration can be added to the right-hand side of Eq.
(38) which can be balanced by the tenacity of particle attachment to
The particle weight in Eq. (38) is constant, whilst the buoyancy determine the upper limit of floatable particle size. For a particle much
supporting the stability of the particle is a function of the polar contact smaller than the relevant bubble it can be shown that the maximum
position. Due to the geometry constraint,θ = α + β, the capillary force particle size that remains attached to a bubble is as follows:
exhibits a maximum at α=β=θ/2. Buoyancy also has a maximum, called
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
the tenacity of particle attachment, which can be described as follows: 3σ ð1− cosθÞ
n o Rp max ¼ : ð43Þ
pffiffiffiffiffiffi 4ρp ðg þ bm Þ
T ¼ maxf F b g ¼ πRp σ ð1− cosθÞ 1 þ 0:016 Bo þ OðBoÞ ð40Þ

pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi For larger particles, the maximum particle size can be obtained by
where Bo=(Rp/L)2 is the Bond number and L ¼ σ =ðδgÞ is the capillary pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
length, which is about 2.7 mm for the air-water surface at room temper- solving the following nonlinear equation: πRp σ ð 1 þ 2A cosθ þ A2 −
ature. The cut-off terms of Eq. (40) are of the order of the Bond number, 4πR 3 ρ
cosθ−AÞ ¼ 3p p ðg þ bm Þ. Table 3 shows the results of the calculation
which is satisfactorily met by the particle size in flotation (Bo≪1). and comparison with flotation experimental data. It should be
The effect of the bubble size on the tenacity of particle attachment recognised that the results in Table 3 are approximate since surfactant
has been investigated by Nguyen (Nguyen, 2003), whom showed that adsorption and desorption in turbulent flow are known to influence
the tenacity of a bubble-particle couplet is described as follows: hydrophobicity and bubble-particle interactions (Omelka et al., 2010),
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
  as well as the particle shape and surface morphology can also be signif-
T ¼ πRp σ 1 þ 2A cosθ þ A2 − cosθ−A þ O Bo3=2 ð41Þ icant but are not considered in the modelling yet.
The calculation of the machine acceleration described by Eq. (42)
requires the bubble size, which is usually a variable in flotation. It can
where A = Bo(1− Rb/L).
be estimated based on the effect of microturbulence on the stability of
Knowing the tenacity of particle attachment then the efficiency of
air bubbles, i.e., the bubble size in the impeller region is also controlled
bubble-particle stability and detachment can be computed and then
by pressure fluctuation of microturbulence (Nguyen and Schulze,
used to predict the maximum size of floatable particle (Pyke et al.,
2004). The pressure tends to break-up the bubble, which is resisted by
2003; Nguyen and Schulze, 2004; Jameson et al., 2007; Goel and
the surface tension force. Eq. (12) for the pressure fluctuation of the
Jameson, 2012). In addition to the particle weight, turbulence has a sig-
inertial subrange is usually used. Balancing the two forces gives the
nificant role in disrupting the bubble-particle aggregates. In particular,
following prediction for the bubble size (Jameson et al., 2007):
in a mechanical flotation cell the region surrounding by the impeller
and stator is highly turbulent and it is here that most of the energy is !
imparted to the flotation slurry to produce air bubbles and promote σ 3=5
Rb max ¼ 3:27 ð44Þ
the bubble-particle interactions. Turbulence in the impeller-stator re- ε i 2=5 δ3=5
gion influences not only the collision but also the detachment. Typically,
the results are described in terms of either “machine acceleration”, bm, where εi is the power per unit mass based on the mass of liquid in the
or the mean energy dissipation rate. Schulze (Schulze, 1983) considered volume swept by the impeller. The machine acceleration appropriate
that air bubbles in the impeller region behaved as if they were at the to the maximum bubble size is obtained by substituting Eq. (44) into
centre of a vortex, so they rotated with the vortex as shown in Fig. 7. Eq. (42), giving:
The rotational velocity was found using isotropic turbulence theory.

Any particle on the surface of the bubble would experience a centrifugal bm ¼ 1:28εi 4=5 δ1=5 =σ 1=5 : ð45Þ
force tending to move the particle away from the bubble. For the
machine acceleration case, Schulze (1982) used the acceleration of Inserting Eq. (45) into Eq. (43), with bm ≫g, yields the following approx-
turbulence in the inertial subrange described by Eq. (15), which gives imate equation for the maximum floatable particle size:
( )1=2
ε2=3 σ 6=5 ð1− cosθÞ
bm ¼ 1:9 1=3
: ð42Þ Rp max ¼ 0:77 : ð46Þ
Rb ρp εi 4=5 δ1=5

Turbulent energy dissipation rate was experimentally correlated Eq. (46) can be tested using the following conditions (Jameson et al.,
with the probability of bubble-particle detachment for the first time 2007):
only recently (Goel and Jameson, 2012). It is noted that in place of the
1) Mechanical flotation cell with power input of 3 kW/m3,
2) Swept volume of the impeller is one-tenth of the volume of the
liquid in the cell, giving a power consumption in the impeller region
of 30 kW/m3,

Table 3
Dependence of the maximum size of floatable sylvinite particles on the acceleration of
turbulent eddies (Nguyen, 2002).

bm/g Rp max (μm)

Eq. (43) Experimental

4.45 708 700


8.75 529 480
Fig. 7. A bubble-particle couplet rotating in a turbulent eddy (left) and the forces acting on
12.10 457 430
the particle (right) (Jameson et al., 2007).
84 A.V. Nguyen et al. / International Journal of Mineral Processing 156 (2016) 75–86

3) Chalcopyrite particles of density 4200 kg/m3, in a pulp of density spheres as shown in Fig. 8. For small angles of truncation, GIC did
1200 kg/m3, and not affect the sphere detachment and hence the classical theories on
4) Surface tension of 60 mN/m, and the contact angle of 60 degrees. the floatability of spheres are valid. For large truncated angles, GIC deter-
mined the tenacity of the particle-meniscus contact and the stability and
Eq. (46) gives a maximum floatable particle size of 500 μm, which is
detachment of floating spheres, and the classical theories were invalid.
within the range of expectations.
The detaching force between an air bubble and a cubical particle has
Knowing the maximum floatable particle size, the efficiency, Es, of
also shown to be influenced by the shape edge of the particle (Gautam
the bubble-particle aggregate stability in the turbulent field can be
and Jameson, 2012).
predicted as follows:

(
) ( ) 6. Conclusions
Rp max 2 0:59σ 6=5 ð1− cosθÞ
Es ¼ 1− exp 1− ¼ 1− exp 1− :ð47Þ
Rp Rp 2 ρp εi 4=5 δ1=5 In this paper the literature on the quantification of the effect of
microturbulence on bubble-particle interaction in flotation has been
reviewed. While the theory on microturbulence of water flow is well
developed, the stochastic description of particle and bubble motions in
5.1. Limits of the current theoretical developments
the turbulent flow field in the mechanical flotation cells is limited. The
deterministic modelling of bubble–particle collision interaction is well
The presented theoretical developments for bubble-particle stability
developed and can be used to approximate the interaction by the
and detachment are derived when the turbulent tensile stresses are
mean (time-averaged) turbulent flow. The effect of microturbulence
dominant. Shear stresses outlined by the transverse acceleration de-
on bubble-particle collision is less quantified despite the literature on
scribed by Eq. (14) can also be significant. If the turbulent shear stresses
particle-particle turbulent collision in gas-particle flow and coagulation
or the vibration of bubble-particle aggregates rising to the pulp-froth
is very rich. The application of the particle-particle collision models for
interface are the dominant forces causing the particle detachment, the
flotation is not straightforward because of the different properties of
theories have to be modified, as shown in the literature (Nguyen and
air bubbles, namely, large size and rise velocity as opposite to settling
Schulze, 2004). Significantly, the current theories are based on constant,
of particles. Further analysis of the particle-particle collision models
single-valued contact angle. Practically, the contact angle is multi-valued
reveals similar mechanisms governing bubble-particle collision in
and varies from the receding (minimum) to advancing (maximum)
flotation because the governing forces are similar. The cylindrical
values. The contact variation known as the contact angle hysteresis is
formulation for modelling turbulent collision is enormous and should
due to the fact that the particle surface is almost always rough and chem-
not be used. Abrahamson's turbulent collision frequency functions
ically heterogeneous; and therefore the resistance to the triple contact
modified to account for the variance of fluctuating velocities for air
line relaxation during the advancing event is different from that during
bubbles and solid particles would be a good approximation for flotation.
the receding event. Since detachment is an advancing event, the advanc-
No quantitative models of the bubble-particle turbulent collision effi-
ing contact angle can be applied to the current theory if the contact angle
ciency relevant for flotation are available. Modelling bubble-particle
is narrowly distributed and contact angle hysteresis is small. Recent
turbulent detachment has been based on balancing the tenacity of
work (Feng and Nguyen, 2016) showed that the pinning of the contact
buoyancy forces and centrifugal force using the mean machine acceler-
line at the sharp edge, known as the Gibbs inequality condition (GIC),
ation estimated from microturbulence theory. There still exist a number
can play a significant role in controlling the stability and detachment of
of deficiencies in modelling bubble-particle turbulent detachment. The
floating spheres. In this study, the spheres were truncated with different
majority of studies only considers the particle turbulent acceleration
angles and the force of pushing the truncated spheres from the interface
in the centrifugal direction but ignore the transverse acceleration of par-
into water was measured. Both the experimental and theoretical results
ticles, which is due to turbulent shear flow. Contact angle required in
confirmed the critical effect of GIC on stability and detachment of floating
quantifying the detachment is not constant as considered in the current
theories, but are multi-valued and can invalidate the theories if there is
a significant contact angle hysteresis. The latest experiments based on
the Gibbs inequality condition show that multi-valued contact angle
can significant affect stability and detachment of floating particles. Fur-
ther research is required for quantifying the effect of microturbulence
on bubble-particle interaction in flotation.

Nomenclature
aλ eddy acceleration
Bll longitudinal correlation function of turbulent acceleration
Bnn transverse correlation function of turbulent acceleration
Bo Bond number
bm machine acceleration
bm⁎ machine acceleration with maximum bubble size
C numerical constant
D structure function of fluctuating velocity
DL Lagrangian structure function of fluid flow
Fig. 8. Effect of pinning of the contact line at the sharp edge, known as the Gibbs inequality
Dll longitudinal structure function of fluctuating velocity
condition, on stability and detachment of floating spheres (Feng and Nguyen, 2016). The
vertical axis shows the supporting force of buoyancy (depicted in Fig. 6) normalised by Dnn transverse structure function of fluctuating velocity
dividing by 2πσRp. During the detachment by advancing of triple contact line from the Dpp structure function of fluctuating pressure
spherical surface to the planar surface of truncation, the advancing contact angle Ec collision efficiency
changes from θ to θ+α0 +π/2, where α0 is the half-filling angle of truncation, leading to Ec,0 collision efficiency as St→ 0
a shape increase in the supporting force. For a large angle of truncation, the increase in
the supporting force exceeds the tenacity of attachment predicted at α = θ/2
Ec,i collision efficiency by interception
(approximately) by the classical theory which otherwise fails to explain the stability of Ec,in collision efficiency by inertia
the floating sphere in this case. Ec,g collision efficiency by gravity
A.V. Nguyen et al. / International Journal of Mineral Processing 156 (2016) 75–86 85

Ec,gi combined collision efficiency by gravity and interception Acknowledgements


Ec,t collision efficiency by turbulence
Es efficiency of the bubble-particle aggregate stability This research is supported under Australian Research Council's
Fb buoyancy force Projects funding schemes (project number DP150100395).
Fg particle weight
f interpolating function
References
fd drag correction factor
g gravity acceleration vector Abrahamson, J., 1975. Collision rates of small particles in a vigorously turbulent fluid.
H meniscus depression Chem. Eng. Sci. 30 (11), 1371–1379.
Anfruns, J.F., Kitchener, J.A., 1977. Rate of capture of small particles in flotation. Trans. Inst.
J particle flux vector Min. Metall. Sect. C 86, 9–15.
L capillary length Batchelor, G.K., 1982. The Theory of Homogenous Turbulence. Cambridge University
mp particle mass Press, Cambridge (197 pp).
Brady, M.R., Telionis, D.P., Vlachos, P.P., Yoon, R.-H., 2006. Evaluation of multiphase flota-
mb bubble mass
tion models in grid turbulence via Particle Image Velocimetry. Int. J. Miner. Process.
Nc bubble-particle collision rate 80, 133–143.
Nci ideal collision rate Dai, Z., Dukhin, S., Fornasiero, D., Ralston, J., 1998. The inertial hydrodynamic interaction
Ncr real collision rate of particles and rising bubbles with mobile surfaces. J. Colloid Interface Sci. 197 (2),
275–292.
np particle number density Dai, Z., Fornasiero, D., Ralston, J., 2000. Particle-bubble collision models — a review. Adv.
p probability distribution function Colloid Interf. Sci. 85 (2–3), 231–256.
R sum of two particle radii Dobby, G.S., Finch, J.A., 1986. A model of particle sliding time for flotation size bubbles.
J. Colloid Interface Sci. 109 (2), 493–498.
Rb bubble radius Dobby, G.S., Finch, J.A., 1987. Particle size dependence in flotation derived from a
Rp particle radius fundamental model of the capture process. Int. J. Miner. Process. 21 (3–4),
Reλ Reynolds number of a turbulent eddy 241–260.
Dukhin, S.S., 1983. Critical value of the Stokes number and the Sutherland formula.
r distance vector between two points Kolloidn. Zh. 45 (2), 207–218.
St Stokes number Feng, D., Nguyen, A.V., 2016. How does the Gibbs inequality condition affect the stability
T particle relaxation time with a “joined mass” and detachment of floating spheres from the free surface of water? Langmuir 32.
http://dx.doi.org/10.1021/acs.langmuir.5b04098.
Tλ oscillation period of an eddy
Flint, L.R., Howarth, W.J., 1971. Collision efficiency of small particles with spherical air
t time bubbles. Chem. Eng. Sci. 26 (8), 1155–1168.
Ub bubble rising velocity Gaudin, A.M., 1957. Flotation. McGraw-Hill, New York.
Gautam, A., Jameson, G.J., 2012. The capillary force between a bubble and a cubical
V particle velocity vector
particle. Miner. Eng. 36–38, 291–299.
Vp particle settling velocity Goel, S., Jameson, G.J., 2012. Detachment of particles from bubbles in an agitated vessel.
W liquid velocity vector Miner. Eng. 36-38, 324–330.
Wλ velocity of a turbulent eddy Guerra, E.A., Schubert, H., 1997. Effects of some important parameters on fine particle
entrainment during flotation in mechanical cells. Proc. XX Int. Miner. Process.
WK Kolmogorov velocity scale Congr., Aachen, Germany, pp. 153–165.
x position vector Jameson, G.J., Nam, S., Young, M.M., 1977. Physical factors affecting recovery rates in
ΔP root mean square of fluctuating pressure difference flotation. Miner. Sci. Eng. 9 (No. 3. July 1977), 103.
Jameson, G.J., Nguyen, A.V., Ata, S., 2007. The flotation of fine and coarse particles. In:
ΔV particle velocity vector relative to bubble Fuerstenau, M.C., Jameson, G.J., Yoon, R.-H. (Eds.), Froth Flotation: A Century of
ΔW square root of longitudinal structure function of fluctuating Innovation. SME, Denver, CO, USA, pp. 329–351.
velocity Jiang, Z.W., Holtham, P.N., 1986. Theoretical model of collision between particles and
bubbles in flotation. Inst. Min. Metall. 95, C187–C194.
King, R.P., 2001. Modeling and Simulation of Mineral Processing Systems. Elsevier.
Greek Kolmogorov, A.N., 1941. Local structure of turbulence in noncompressible fluid with very
high Reynolds number. Doklady Acad. Sci. USSR 30, 168–181.
Kostoglou, M., Karapantsios, T.D., Matis, K.A., 2006. Modeling local flotation frequency in a
α central angle
turbulent flow field. Adv. Colloid Interf. Sci. 122 (1), 79–91.
β collision frequency function Kruis, F., Kusters, K., 1997. The collision rate of particles in turbulent flow. Chem. Eng.
β′ collision frequency function due to fluctuating motions Commun. 158 (1), 201–230.
Liepe, F., Moeckel, H.O., 1976. Studies of the combination of substances in liquid phase.
δ liquid density
Part 6: the influence of the turbulence on the mass transfer of suspended particles.
δij Kronecker delta Chem. Tech. 28 (4), 205–209.
ε turbulent kinetic energy dissipation rate Lynch, A.J., Harbort, G.J., Nelson, M.G., 2010. History of Flotation. The Australasian Institute
εi power per unit mass of Mining and Metallurgy, Burwood, Vic, Australia (348 pp).
Meyer, C.J., Deglon, D.A., 2011. Particle collision modelling — a review. Miner. Eng. 24,
θ contact angle 719–730.
Λ macro-turbulence length scale Nguyen, A.V., 2002. New method and equations for determining attachment tenacity and
λ length scale of a turbulent eddy particle size limit in flotation. Int. J. Miner. Process. 68, 167–182.
Nguyen, A.V., 2003. New method and equations for determining attachment tenacity and
λD Taylor microscale of fluid acceleration particle size limit in flotation. Int. J. Miner. Process. 68 (1), 167–182.
λk Kolmogorov turbulent length scale Nguyen, A.V., 2011. Particle-bubble interaction in flotation. In: Miller, R., Liggieri, L. (Eds.),
μ liquid dynamic viscosity Bubble and Drop InterfacesProgress in Colloid and Interface Science. Brill, Leiden, The
Netherland.
ν liquid kinematic viscosity Nguyen, A.V., Schulze, H.J., 2004. Colloidal Science of Flotation. Marcel Dekker, New York
ξ surface vorticity (840 pp).
ρp particle density Nguyen, A.V., George, P., Jameson, G.J., 2006. Demonstration of a minimum in the
recovery of nanoparticles by flotation: theory and experiment. Chem. Eng. Sci. 61,
ρb bubble density
2494–2509.
σ standard deviation of a distribution or surface tension Omelka, B., Schreithofer, N., Heiskanen, K., 2010. Effect of hydrophobicity and frother
τ1 , τ2 particle relaxation times concentration on bubble-particle interactions in turbulent flow. Proceedings XXV
International Mineral Processing Congress, Santiago, Chile, pp. 2205–2214.
τmax maximal laminar shear stress
Panchev, S., 1971. Random Functions and Turbulence. Pergamon Press, Oxford.
ϕ turbulent variable Phan, C.M., Nguyen, A.V., Miller, J.D., Evans, G.M., Jameson, G.J., 2003. Investigations of
ϕ mean component of a turbulent variable bubble–particle interactions. Int. J. Miner. Process. 72 (1), 239–254.
ϕ′ fluctuating component of a turbulent variable Pyke, B., Fornasiero, D., Ralston, J., 2003. Bubble particle heterocoagulation under
turbulent conditions. J. Colloid Interface Sci. 265 (1), 141–151.
φc collision angle Ralston, J., Dukhin, S.S., Mishchuk, N.A., 1999. Inertial hydrodynamic particle-bubble
ψ stream function interaction in flotation. Int. J. Miner. Process. 56 (1–4), 207–256.
86 A.V. Nguyen et al. / International Journal of Mineral Processing 156 (2016) 75–86

Reay, D., Ratcliff, G.A., 1973. Removal of fine particles from water by dispersed air Schulze, H.J., 1983. Physicochemical Elementary Processes in Flotation. Vol. 1983. Elsevier
flotation. Effects of bubble size and particle size on collection efficiency. Can. Science Publishers, p. 348.
J. Chem. Eng. 51 (2), 178–185. Schulze, H.J., 1989. Hydrodynamics of bubble-mineral particle collisions. Miner. Process.
Reay, D., Ratcliff, G.A., 1975. Experimental testing of the hydrodynamic collision model of Extr. Metall. Rev. 5, 43–76.
fine particle flotation. Can. J. Chem. Eng. 53 (5), 481–486. Sutherland, K., 1948. The physical chemistry of flotation XI. Kinetics of the flotation
Saffman, P., Turner, J., 1956. On the collision of drops in turbulent clouds. J. Fluid Mech. 1 process. J. Phys. Chem. 52, 394–425.
(01), 16–30. Wang, L.-P., Wexler, A.S., Zhou, Y., 1998. Statistical mechanical descriptions of turbulent
Schubert, H., 1999. On the turbulence-controlled microprocesses in flotation machines. coagulation. Phys. Fluids 10, 2647–2651.
Int. J. Miner. Process. 56 (1), 257–276. Weber, M.E., Paddock, D., 1983. Interceptional and gravitational collision efficiencies for
Schubert, H., Bischofberger, C., 1978. On the hydrodynamics of flotation machines. Int. single collectors at intermediate Reynolds numbers. J. Colloid Interface Sci. 94 (2),
J. Miner. Process. 5 (2), 131–142. 328–335.
Schubert, H., Bischofberger, C., 1998. On the microprocesses air dispersion and particle- Williams, J., Crane, R., 1983. Particle collision rate in turbulent flow. Int. J. Multiphase Flow
bubble attachment in flotation machines as well as consequences for the scale-up 9 (4), 421–435.
of macroprocesses. Int. J. Miner. Process. 52 (4), 245–259. Yoon, R.H., Luttrell, G.H., 1989. The effect of bubble size on fine particle flotation. Miner.
Schubert, H., Heidenreich, E., Liepe, F., Neesse, T., 1990. Mechanische Verfahrenstechnik. Process. Extr. Metall. Rev. 5, 101–122.
Deutscher Verlag für Grundstoffindustrie, Leipzig (407 pp). Zaichik, L.I., Alipchenkov, V.M., Sinaiski, E.G., 2008. Particles in Turbulent Flows. Wiley-
Schulze, H.J., 1982. Dimensionless number and approximate calculation of the upper par- VCH Verlag GmbH, Weinheim, Germany.
ticle size of floatability in flotation machines. Int. J. Miner. Process. 9 (4), 321–328.

You might also like