Recovery of Fine and Ultrafine Mineral Particles by Electroflotation - A Review

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Mineral Processing and Extractive Metallurgy Review

An International Journal

ISSN: 0882-7508 (Print) 1547-7401 (Online) Journal homepage: http://www.tandfonline.com/loi/gmpr20

Recovery of fine and ultrafine mineral particles by


electroflotation – A review

Bogale Tadesse, Boris Albijanic, Fidele Makuei & Richard Browner

To cite this article: Bogale Tadesse, Boris Albijanic, Fidele Makuei & Richard Browner (2018):
Recovery of fine and ultrafine mineral particles by electroflotation – A review, Mineral Processing
and Extractive Metallurgy Review, DOI: 10.1080/08827508.2018.1497627

To link to this article: https://doi.org/10.1080/08827508.2018.1497627

Published online: 19 Jul 2018.

Submit your article to this journal

Article views: 81

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=gmpr20
MINERAL PROCESSING AND EXTRACTIVE METALLURGY REVIEW
https://doi.org/10.1080/08827508.2018.1497627

Recovery of fine and ultrafine mineral particles by electroflotation – A review


Bogale Tadesse, Boris Albijanic, Fidele Makuei, and Richard Browner
Department of Mining Engineering and Metallurgical Engineering, Western Australian School of Mines, Curtin University, Kalgoorlie, Australia

ABSTRACT KEYWORDS
Fine and ultrafine particles are successfully recovered by flotation when fine bubbles are present. Ultrafine particles;
Electroflotation is a technique in which fine bubbles are generated by the electrolysis of water. This electroflotation; electrolytic
article reviews the experimental studies into mineral recovery by electroflotation and the potential bubbles
application of electrolytic bubbles in the recovery of fine and ultrafine mineral particles. The literature
reveals that electroflotation resulted in better recoveries of ultrafine particles (e.g. dolomite, magnesite,
and pyrite) as compared with dispersed-air flotation because electrolytic bubbles are smaller in size and
are more active than those generated during dispersed-air flotation.

1. Introduction bubbles with a mean diameter from 17 to 105 μm based on


the conditions of electrolysis (Han et al. 2002; Jiménez et al.
Flotation method involves the introduction of air bubbles into
2010). The technique is particularly attractive because bubbles
ore slurries in which the target particles selectively attach to
sizes can be controlled by varying the current density, pulp
air bubbles and rise to the surface, overflowing the flotation
pH, and electrode materials.
cell into the flotation launder. This method works well for
particles of a limited size range, beyond which its efficiency The application of flotation for the recovery of valuable
reduces substantially. More precisely, particles finer than minerals originated from Broken Hill, Australia in 1904
approximately 10 μm are in general not effectively and selec- (Elmore 1905). Some reports suggest that the technique of
tively separated by traditional froth flotation because they electroflotation originated mainly in the USSR for the treat-
tend to follow liquid streamlines around bubble due to their ment of waste water and the clarification of grape liquor
low inertia, reducing the probability of collision between (Matis and Peleka 2010). Commercial applications of elec-
particles and bubbles (Zhou et al. 1997; Dai et al. 1998). A troflotation are reported in silk and meat industries, oil
number of factors such as contact angle, surface roughness, refining, leather industries and steel plants mostly in the
hydrophobicity, zeta potential, reagent type and pH are former USSR and a few in Germany, Japan, and the UK
known to affect flotation performance because of their impact (Bhaskar Raju and Khangaonkar 1984a; Matis and Gallios
on the adsorption of collectors on mineral particles. However, 1986). Other small scale studies include the electroflotation
in the case of ultrafine particles, due to their higher specific of pyrite (Kyrdos et al., 1994), chalcopyrite (Bhaskar Raju
surface area and higher surface energy, the adsorption of and Khangaonkar 1982), sphalerite (Llerena et al. 1996),
collectors becomes non-specific. That is, fine and ultrafine dolomite and magnesite (Matis et al. 1993), and silica
particles behave differently from coarse particles during flota- (Sarkar et al. 2011), removal of heavy and toxic metals
tion. The challenges encountered during flotation of fine and from wastewater (Zouboulis and Matis 1989, 2012; Il’in
ultrafine particles by mechanical cells are well documented et al. 2014; Oliveira et al., 2015), and the separation of oil
(Gaudin et al. 1931, 1942; Flint and Howarth 1971; Trahar and low-density suspended solids from various wastewaters
and Warren 1976; Fuerstenau 1980; Yoon and Luttrell 1989; (Alexandrova et al. 1994; Kubritskaya et al. 2000; Chen
Miettinen et al. 2010). Thus, fine and ultrafine particle separa- 2004; Kyzas and Matis 2016).
tion by flotation is now considered as a major challenge in The focus of this work is on the physics governing the
mineral recovery and extensive research is necessary to generation of electrolytic bubbles and the potential applica-
improve flotation performance (Ahmadi et al., 2014). tion of electrolytic bubbles to recovery fine particles by elec-
Given that the probability of collection of fine particles troflotation. The modeling of electroflotation process will also
increases with decreasing bubble size (Dai et al. 2000), various be reviewed.
techniques have been developed for the generation of fine The article is organized as follows. The mechanisms of
bubbles for flotation separation. In electroflotation, bubbles bubble formation in electroflotation will be briefly reviewed
are generated through the electrolytic breakdown of water in the next section. This is followed by an overview of factors
which leads to the formation of fine hydrogen and oxygen affecting bubble size during electroflotation. Advantages and

CONTACT Bogale Tadesse Bogale.tadesse@curtin.edu.au Department of Mining Engineering and Metallurgical Engineering, Western Australian School of
Mines, Curtin University, Egan St, Kalgoorlie, WA 6430, Australia
Color versions of one or more of the figures in the article can be found online at www.tandfonline.com/gmpr.
© 2018 Taylor & Francis Group, LLC
2 B. TADESSE ET AL.

disadvantages of the use of electroflotation in fine and ultra- 2.1 Bubble Nucleation on an Electrode
fine particle recovery will be described in Section 4. The
The nucleation of a bubble on an electrode surface occurs
modeling of electroflotation process will be reviewed in the
when the bubble attains a radius that is larger than the critical
last section.
bubble radius (r ), which is defined by the Gibbs-Thompson
Before reviewing electroflotation, it is important to highlight
equation as:
that the classification of particles into coarse, fine, and ultrafine.
A particle size classification for gravity separation has been 2σΩ
r ¼ (4)
proposed elsewhere (Sivamohan and Forssberg 1985). kTlnðp=p0 Þ
However, in this article in relation to electroflotation, coarse
particles range in size from 38 to 250 µm, fine particles are those where σ, Ω, k, and T are the interfacial tension, the molecular
in the range between 10 and 38 µm while ultrafine particles are volume, the Boltzmann’s constant, and the temperature,
those in the range between 1 and 10 µm in diameter. respectively; p0 is the vapor pressure of a flat sample of
infinite radius, while p is the vapor pressure of a sub-critical
nucleus of radius, r . As can be seen in Equation (4), the
increase in p leads to decrease in r .
2. Mechanisms of Bubble Formation in In electroflotation, nucleation of bubbles occurs on solid
Electroflotation electrodes. The rate of heterogeneous bubble nucleation, J, is
given by (Lubetkin 1995):
Hydrogen and oxygen bubbles are generated during water
 
electrolysis (Matis et al. 1993). More precisely, hydrogen is Φ0 ðθÞ ΔG
released at the cathode surface while oxygen is released at the J ¼ Cexp (5)
kT
anode surface as shown in Equations (1) and (2).
where C is pre-exponential factor, Φ0 ðθÞ is the fractional
reduction in the volume of the critical sized bubble when it
Anode : 2H2 OðlÞ ! O2ðgÞ þ 4H þ ðaqÞ þ 4e (1)
forms on a surface:

Cathode : 4H þ ðaqÞ þ 4e ! 2H2 ðgÞ (2) Φ0 ðθÞ ¼ ð1 þ cosθÞ2 ð2  cosθÞ=4 (6)

The free energy barrier to the phase change, ΔG , is given by:
The cell voltage between two electrodes during water electro-  2
lysis is the summation of the equilibrium potential difference, ΔG ¼ 16πσ 3 Ω2 =3 pb  p (7)
the anodic and cathodic overpotentials, and ohmic potential and pb is the pressure inside the bubble.
drop of the aqueous solution. Bubble nucleation rate depends on electrode surface inhomo-
The standard free energy of reaction, ΔG0 ; can be geneity, the diameter of an electrode wire, and current density
expressed in terms of the thermodynamic potential (E0 ) of (Glas and Westwater 1964; Janssen and Hoogland 1973). The
water decomposition as: inhomogeneity of the electrode surface (e.g. cracks and
scratches) results in the increase in bubble nucleation rate due
ΔG0 to atomic ledges and thus higher energy points on the electrode
E0 ¼ (3)
nF surface. The smaller diameter of the electrode wire also leads to
higher bubble nucleation rate probably because bubbles
where n is the number of exchanged electrons and F is detached from wire with a small diameter have a smaller size
Faraday’s constant (96,485 C/mol). than those detached from wire with a larger diameter (Sarkar
Since the decomposition of water at standard temperature et al. 2010). Thus, the detachment rate of bubbles from the wire
and pressure requires a theoretical minimum of 237.13 kJ/mol with smaller diameter results in more nucleation sites on the
of electrical energy input (Chen and Chen 2010), it follows wire surface. Finally, the increase in current density leads to
from Equation (3) that the minimum voltage necessary for higher bubble nucleation rate probably because the generation
electrolysis is about 1.23 V. Although the equilibrium poten- of molecules is more pronounced at a higher current density (i.e.
tial difference required for the decomposition of pure water the concentration of dissolved gas increases).
into hydrogen and oxygen −1.23 V, a potential difference of
> 2 V is often required in practice due to the contributions of
overpotantial and ohmic potential drop (Chen et al. 2002). 2.2 Bubble Growth on an Electrode
Hydrogen and oxygen molecules generated by electrolysis Bubbles can grow if bubble size is higher than the critical
forms bubbles but can also dissolve into the bulk solution by bubble size. Another reason for bubble growth is coalescence
convective diffusion. For example, Müller et al. (1989) found with other bubbles on electrode surfaces (Sides and Tobias
that at a current density of 3000 A/m2, up to 85% of hydrogen 1985). A number of factors influence the bubble growth rate
generated forms hydrogen gas bubbles. The mechanism of including the molecular diffusion of gas to the interface of the
formation of bubbles at electrode surfaces involves three bubble, inertia, viscosity, and surface tension. Bubble growth
sub-processes: bubble nucleation, bubble growth, and bubble kinetics can be described using the power law model
detachment from the electrode. Bubble sizes also increase due (Brandon and Kelsall 1985):
to the diffusion of gas through bubble surfaces. These sub-
processes are discussed below. rðtÞ ¼ η0 t l (8)
MINERAL PROCESSING AND EXTRACTIVE METALLURGY REVIEW 3

where rðtÞ is the bubble radius at time t after nucleation, η0 surface. Bubble detachment is dependent on the balance of
is the growth coefficient, t is time and l represents the time adhesion and detachment forces acting on the bubble (Jones,
coefficient. It is important to note that the higher the time 1999a). At equilibrium,
coefficient, the faster the bubble growth. It was proposed
(Brandon and Kelsall 1985) that there are three regions of Fd þ Fs ¼ Fi þ Fp þ Fb (9)
bubble growth based on bubble growth time i.e. Region I, Where Fd is the drag force, Fs is the surface tension force,
Region II, and Region III as shown in Figure 1. Fi is the inertial force, Fp is the pressure force, and Fb is the
Region I (t < 10 ms): The bubble growth is the fastest buoyancy. At relatively low growth rates, Equation (9) can be
during the initial period (i.e. where the value of l is between written as:
0.6 and 0.8 in Figure 1) due to the high local saturation of gas
forming molecules. The lower the surface tension of liquid, Fs ¼ Fp þ Fb (10)
the faster the bubble growth rate. In other words, the decrease
Thus, bubbles separate from an electrode when the detach-
in surface tension of liquid increases the internal bubble
ment force (i.e. the sum of the buoyancy force and the
pressure and thus the bubble growth rate rises.
pressure force inside the bubble) is higher than the attach-
Region II (10 < t < 100 ms): When time is between 10
ment force between the bubble and the electrode (i.e. surface
and 100 ms, the bubble growth rate decreases dramatically
tension force). It should be noted that although there is no
(i.e. l = 0.5) due to the reduced local saturation of gas
experimental evidence showing the importance of the electro-
forming molecules. Thus, the controlling step of bubble
static forces during bubble detachment from an electrode,
growth is the diffusion of dissolved molecules to the bubble
these forces may also play a role in the bubble detachment
surface.
process (Brandon and Kelsall 1985; Lubetkin 1995).
Region III (t > 100 ms): When time is longer than 100 ms,
By using the scheme shown in Figure 2 and balancing the
the bubble growth rate is more reduced than that in Region II
attachment force with the detachment forces, the radius of the
(i.e. l = 0.33). In this region, gas penetrates through the bubble
detaching bubble from an electrode can be determined as
on electrode surfaces while diffusion of gas to the bulk solu-
follows (Sarkar et al. 2010):
tion is negligible.
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
The growth coefficient (η0 ) for hydrogen and oxygen u 24σsin2 θ
u
evolution was found to be dependent on current density, db;d ¼ t   (11)
faradaic charge and electrode rotation speed but indepen- g ρf  ρG ð2 þ 3cosθ  cos3 θÞ
dent on solution pH, electrode material (Pt or Ag) and the and
addition of sodium dodecyl sulphate or dodecyl-trimethyl-
ω
ammonium bromide (Brandon and Kelsall 1985; Brandon θ¼   (12)
et al. 1985). 1 þ sin1 db;d =de
where de is the diameter of the electrode, ω is the liquid
surface tension, db;d is the diameter of the detached bubble,
2.3 Detachment of Bubble from an Electrode Surface θ is the contact angle, ρf and ρG are the density of the liquid
Following the aforementioned nucleation and growth pro- and gas, respectively. The diameter of the detached bubble can
cesses, the bubble eventually detaches from the electrode be calculated by combining Equation (11) with Equation (12).

1.0

Region I
0.8

Region II
Time Coefficient

0.6
Region III

0.4

0.2

0.0
0.01 0.1 1
Time (s)
Figure 2. Geometry of bubble on curved electrode surface which is used to
Figure 1. Change in time coefficient (l) during hydrogen bubble growth on a calculate bubble detachment diameter where h is the height from the top of the
25 µm diameter Pt electrode at pH 1.4 (Brandon and Kelsall 1985). detaching bubble to the electrode surface (Sarkar et al. 2010).
4 B. TADESSE ET AL.

300 2.4 Diffusion of Gas Through Bubble Surface


H2 (-2.2 V, 10 µA) Some fraction of the gas generated during electrolysis can
250 H2 (-1.08 V, 1 µA) diffuse into bubbles separating from the electrode surface,
Bubble detachment diameter (µm)

O2 (3.2 V, 10 µA) causing an increase in bubble sizes. Once the bubble is


O2 (2.1 V, 1 µA) detached from the electrode surface it grows rapidly as it
200
moves through the bulk solution (Sarkar et al. 2010). This
growth was attributed to a number of factors including hydro-
150 static pressure variation and diffusion of hydrogen, water
vapor and other dissolved gases into the bubble.
100 The diffusion rate (n, mol/s) can be described as follows
(Jones et al. 1999b):

50 dn
¼ Sb kG ðCb  Csat Þ (13)
dt
0 where Sb (m2) represents the bubble surface area, kG is mass
0 2 4 6 8 10 12 14 transfer coefficient; Cb (mol/m3) and Csat (mol/m3) are the
pH molar concentration in the bulk liquid and the saturated
Figure 3. Effect of pH on bubble detachment diameter of hydrogen and oxygen
molar concentration on the liquid–gas interface, respectively.
on a 10 µm diameter platinum electrode (Brandon and Kelsall 1985). The mass transfer coefficient can be determined using
Equation. (14) (Vogt 1984):
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Experimental results show that in surfactant-free solu- 2lnð2ÞD  ρG
tions the detachment diameter of hydrogen bubbles was kG ¼ 0:5  (14)
tMG ðCb  Csat Þ
large at low pH and small at high pH while the opposite
trend was observed for oxygen bubbles as shown in where D (m2/s) represents the mass diffusivity, ρG is the
Figure 3 (Brandon and Kelsall 1985). The reason behind density of gas (kg/m3) and MG (kg/kmol) is the molecular
these observations was explained by the electrostatic inter- mass of the gas molecules.
action between the charged electrodes from which the bub- The gas concentration in the bulk liquid changes with time and
bles are generated and the charges on the bubbles. position in relation to an electrode. For example, in the case of a
Electrolytically generated hydrogen and oxygen bubbles wire cylindrical electrode, gas concentration (including dissolved
have been shown to possess a net charge except at the molecules) in the bulk liquid is given by (Sarkar et al. 2010):
point of the zero charge at pH 2–3 (Brandon et al. 1985).   !  
At pH < 2, bubbles are positively charged while they are dw ðr2 þ ðdw =2Þ2 rdw
Cðr; t Þ ¼ N exp I0 (15)
negatively charged at pH > 3. Thus, at pH < 2 the electro- 4Dt 4Dt 4Dt
static interaction between a negatively charged cathode and
the positively charged hydrogen bubbles results in large where Cðr; t Þ (mol/m3) is the gas concentration at radial
hydrogen bubble detachment diameters, while the opposite position, r, from the cylindrical electrode, t is diffusion time,
is true at pH > 3 (Figure 3). By contrast, the interaction dw is diameter of the cylindrical electrode, N (mol/m2) is
between a positively charged anode and positively charged strength, and I0 is the modified Bessel function of the first
oxygen bubbles led to smaller departure diameters at pH kind of order zero.
< 2. Larger oxygen detachment diameters are expected from
the interaction between the positively charged anode and
the negatively charged oxygen bubbles at pH > 3 as seen in
3. Bubble Size in Electroflotation
Figure 3 (Brandon and Kelsall 1985; Brandon et al. 1985).
However, the influence of pH on departure diameter of Bubble size measurements in electroflotation have been con-
both hydrogen and oxygen bubbles is less pronounced in ducted mainly using photographic techniques (Glas and
the presence of surfactants such as sodium dodecyl sulphate Westwater 1964; Brandon and Kelsall 1985). These measure-
or dodecyl-trimethyl-ammonium bromide. Departure dia- ments were performed for hydrogen bubbles, oxygen bubbles
meter decreases with increasing concentration of surfac- or bubbles generated during electroflotation (i.e. hydrogen
tants. It was suggested that the adsorption of surfactants and oxygen bubbles without distinguishing among them).
at the gas-liquid interface decrease the electrostatic attrac- Given that researchers measured bubble sizes using various
tion between bubble and electrode which reduces the effect bubble generation techniques, measurement positions, and
of pH on the departure diameter (Brandon and Kelsall electrode materials, differing observations for bubble sizes
1985; Brandon et al. 1985). have been reported. Figure 4 shows various factors affecting
Sarkar et al. (2010) reported that the detachment diameter of electrolytically generated bubble diameter (Venkatachalam
hydrogen bubbles was not dependent on current density when 1992). The main factors that affect bubble sizes such as pulp
the current density is between 150 and 350 A/m2 while the pH, current density, and type of the electrode (i.e. electrode
detachment diameter increased with increasing Pt wire diameter. material, geometry, and roughness) will be discussed below.
MINERAL PROCESSING AND EXTRACTIVE METALLURGY REVIEW 5

Figure 4. Factors affecting electrolytically generated bubble size (Venkatachalam 1992).

3.1 Effect of Pulp pH 100


Llerena et al., 1996
The size of electrolytically generated hydrogen bubbles as a Glembotsky et al., 1975
Brandon and Kelsall, 1985
function of pH has been investigated by Brandon and 80 Jimenez et al., 2010
Kelsall (1985) who found that hydrogen bubble size is
larger in acidic solutions due to the higher amount of
Bubble Diameter (µm)

hydrogen ions in these solutions. By contrast, smaller oxy- 60


gen bubbles are produced in acidic media while larger
oxygen bubbles are generated in higher pH conditions
(Glembotsky et al. 1975; Brandon and Kelsall 1985). Other 40
studies suggest that strongly acidic (pH 2) and strongly
alkaline (pH 14) media lead to the formation of larger
bubbles while a larger number of smaller sized bubbles 20
were obtained at neutral pH (Jiménez et al. 2010). By
contrast, other authors have not observed any relationship
between hydrogen bubble size and pulp pH (Llerena et al. 0
1996). The dependence of hydrogen and oxygen bubble size 0 2 4 6 8 10 12 14
on pulp pH under various experimental conditions is sum- pH
marized in Figures 5–6.
Figure 5. Variation of electrolytically generated hydrogen bubble size (diameter)
as a function of electrolyte pH as reported in various studies (Glembotsky et al.
1975; Brandon and Kelsall 1985; Jiménez et al. 2010) as shown by symbols.
3.2 Effect of Current Density
The amount of hydrogen and oxygen gas generated from between current density and bubble size. The influence of
electrodes is directly proportional to the current density. For current density on bubble size under various experimental
that reason, the rise in current density results in the increase conditions is summarized in Figure 7.
in bubble sizes probably due to more pronounced bubble
coalescence (Landolt et al. 1970; Ben Mansour et al. 2007).
Jiménez et al. (2010) found that a large number of small
3.3 Effect of Electrode Material, Geometry and
bubbles were generated at higher current densities while a
Roughness
small number of large bubbles were observed at lower current
densities. Other studies have suggested no clear linkage The selection of electrode material affects bubble size of both
between bubble size and current density (Burns et al. 1997). hydrogen and oxygen bubbles. For example, iron electrodes
It can be concluded that while current density might play produced smaller hydrogen bubble sizes than platinum elec-
some role in controlling bubble sizes and bubble concentra- trodes (Glembotsky et al. 1975). The reason is probably that the
tion, it is necessary to take into account the effect of other iron electrode suppressed the growth and coalescence of hydro-
factors such as electrode material and geometry, and pulp gen bubbles while the opposite is true in the case of the plati-
chemical environment in order to establish the relationship num electrode. It is worth mentioning that the effect of
6 B. TADESSE ET AL.

100 Table 1. Bubble sizes generated on various electrode materials during electro-
Glembotsky et al., 1975 lysis (Modified from Chen and Chen 2010).
Brandon and Kelsall, 1985 Size range (µm) Electrode Reference
80 H2 15–65 Al Han et al. (2002)
H2 19–38 Graphite Burns et al. (1997)
H2 20–89 Steel wire screen Ren et al. (2014a)
Bubble Diameter (µm)

H2 15–23 Pt Sarkar et al. (2010)


60 H2 68–105 Ti Jiménez et al. (2010)
62–88 Stainless steel
H2 25–55 Pt Glembotsky et al. (1975)
17–40 Fe
25–35 C
40
O2 25–55 Pt
H2 22–34 Stainless steel plate Ketkar et al. (1991)
28–39 200 mesh
32–45 100 mesh
20 37–49 60 mesh
O2 42–48 Pt plate
38–50 200 mesh
O2 17–38 Graphite Chen et al. (2002)
0 O2 49–58 Pb-Ca-Sn alloy Al Shakarji et al. (2011)
2 4 6 8 10 12 14
pH

Figure 6. Variation of electrolytically generated oxygen bubble size (diameter) of mesh electrodes resulted in large bubbles (Lumanauw
as a function of electrolyte pH as reported in various studies. Adapted from 2000), while on the contrary, increased roughness on plate
(Brandon and Kelsall 1985; Glembotsky et al. 1975) as shown by symbols. electrodes led to the generation of smaller hydrogen bubbles
probably due to a larger number of nucleation sites (Jiménez
et al. 2010).
160

Burns et al. (1997)


140 4. Potential Application of Electrolytic Bubbles in
Jiménez et al. (2010)
Sarkar et al. (2010) Fine Particle Separation
Mean Bubble Diameter (µm)

120 Ben Mansour et al. (2007)


Electroflotation experiments are commonly performed using a
100 modified Hallimond tube apparatus (Glembotsky et al. 1975;
Matis et al. 1993; Llerena et al. 1996) shown in Figure 8. The
80 apparatus has an electrolysis chamber which is located at the
bottom of a flotation chamber. In the electrolysis chamber, a
60 pair of electrodes is mounted one on each side of a permeable
membrane with the electrodes connected to a DC power
40
supply for the generation of hydrogen and oxygen gas bub-
20
bles. Such electrode arrangement allows for one type of bub-
bles (hydrogen or oxygen bubbles) to be used for flotation
0 while another type of bubbles can be removed from the
0 13 26 39 electrolysis chamber together with the circulating electrolyte.
2
Current Density (mA/cm ) Electrolytic bubbles have different properties than bubbles
produced during dispersed-air flotation. It should be noted
Figure 7. Effect of current density on the size of bubbles generated by electro- that dispersed-air flotation can be performed using Hallimond
lysis. Adapted from (Burns et al. 1997; Ben Mansour et al. 2007; Jiménez et al.
2010; Sarkar et al. 2010). The symbols represent various references. tubes, mechanically agitated cells, or columns. This section
discusses advantages and disadvantages of using electrolytic
bubbles to recover fine mineral particles.
electrode material on the size of hydrogen and oxygen bubbles
should be considered in conjunction with other variables such
4.1 Advantages of Electroflotation
as electrolyte pH and pulp conditions, current density and the
bubble measurement techniques. A concise review of the influ- 4.1.1 Improved Recovery for Fine and Ultrafine Particles
ence of electrode material on bubble sizes is given in Table 1. Electroflotation has some advantages over dispersed-air flota-
In addition to the type of electrode materials, electrode tion. The most important advantage of electroflotation is the
geometry and roughness also influence bubble sizes generated generation of significantly smaller bubbles than that produced
during electrolysis. For example, Ni mesh electrodes produced using dispersed-air flotation methods (see Table 2). Fine bub-
smaller bubbles than Ni plate electrodes (Lumanauw 2000). In bles, generated during electroflotation, have lower buoyancy
the case of mesh electrode, the larger the diameter of the wires and thus their residence time is longer, increasing the prob-
of the cathode, the bigger the generated bubbles. The reason is ability of collision between particles and bubbles. The increase
that two wires with large diameter have less overlapping sites in the probability of collision particularly improves electro-
than those with a smaller diameter. Additionally, wires with flotation recovery of fine particles. The reason is that fine
larger diameter have smaller curvatures. The higher roughness particles have very low settling rates and their small mass
MINERAL PROCESSING AND EXTRACTIVE METALLURGY REVIEW 7

Figure 8. Schematic diagram of a modified Hallimond tube electroflotation cell. 1, flotation chamber; 2, electrolysis chamber; 3, electrolyte inlet; 4, electrolyte outlet;
5, electrodes; 6, diaphragm; 7, float collection tube.

Table 2. Comparison of bubble size generated in various flotation cells (Sarkar the limited solubility of the fatty acid collector in acidic pH.
2012).
Fatty acid collectors form colloidal precipitate at low pH
Cell type Bubble diameter (µm)
(Joseph-Soly et al. 2015) which may lead to poor collector
Mechanically agitated cell 500–2900
Column flotation 600–1500 performance and thus reduced flotation recovery. In the case
Dissolved air flotation 10–300 of magnesite, electroflotation recoveries were about three
Electroflotation: hydrogen bubble 5–65 times higher than that of dispersed-air flotation. The poor
Electroflotation: oxygen bubble 15–95
floatability of magnesite using dispersed-air flotation method
might be attributed to the presence of ultrafine particles on
prevents detachment of particles from a bubble. For example, the surface of magnesite particles (Matis et al. 1993). That is,
Sarkar et al. (2011) found that the maximum recovery of silica the stirring action of impellers during the conditioning stage
varied from 90 – 96% during electroflotation using Cetyl of dispersed-air flotation promoted coating of magnesite par-
trimethyl ammonium bromide collector at pH 10 for particles ticles with magnesite ultrafine particles.
in size range of 5.3 – 12.3 µm diameter. By comparison, Llerena et al. (1996) investigated the flotation behavior of
flotation recovery of silica during flotation using Rushton sphalerite fine particles (< 25 µm in diameter) using hydrogen
turbine mechanically agitated cell in the presence of tri- and oxygen bubbles generated by sparging and electrolytically.
methylchlorosilane collector, and at pH 6 was around 60 – It was reported that flotation recovery obtained by the elec-
70% for −10 µm size fraction particles (Shi and Fornasiero trolytic bubbles was two times higher than that obtained by
2009). The high electroflotation recovery of silica for 5.3 – sparging bubbles under the same rate of gas production. The
12.3 µm particle size range was attributed to an optimisation differences in flotation recovery were attributed to the differ-
of collision and attachment efficiencies at this particle size ences in the size of gas bubbles used in the two flotation
thereby leading to high collection efficiency as observed by experiments; hydrogen and oxygen bubbles sizes generated
Dobby and Finch (Dobby and Finch 1987). A comparison of by sparging and electrolytically were 200 µm and 16 µm in
flotation recovery of various minerals using electroflotation, diameter, respectively (Llerena et al. 1996).
dissolved-air flotation and mechanical flotation in the fine According to Bhaskar Raju and Khangaonkar (1982), the
and ultrafine particle size range is given in Table 3. flotation of chalcopyrite ultrafine particles (average size of
Although the flotation experiments were conducted at differ- 4.42 µm) resulted in a recovery of up to 90% when electro-
ent conditions (such as pH, reagents used, pulp density, lytically generated gas bubbles were used. In comparison,
bubble sizes etc.), it is believed that Table 3 will provide a under the same conditions, the chalcopyrite recovery was
general comparison among these techniques. 30–33% when oxygen bubbles were generated by sparging. It
Electroflotation resulted in better recoveries for both dolo- was suggested that finer bubble sizes in electroflotation made
mite and magnesite when compared with dispersed-air flota- the process more effective for ultrafine particles than those
tion under similar conditions (i.e. particle size fraction generated by sparging even when a smaller gas flow rate was
−45 + 10 µm, 20 mg/L fatty acid collector, and same pulp used (Bhaskar Raju and Khangaonkar 1982).
pH) as seen in Figure 9a–b (Matis et al. 1993). Figure 9a A comparison between electroflotation and dispersed-air
shows lower dolomite electroflotation recovery at pH 6 com- flotation was also reported for the flotation of pyrite (25–
pared with higher pH values. This observation is attributed to 45 µm) (Kydros et al. 1994). In the presence of 15 ppm
8 B. TADESSE ET AL.

Table 3. Comparison of flotation recovery between electroflotation, dissolved-air Figure 10a is considered to be typical of pyrite as confirmed
flotation and mechanical flotation for some selected minerals in the fine and
ultrafine particle size range.
in previous studies (Fuerstenau et al. 1968; Fuerstenau and
Mishra 1981). By contrast, in the presence of 100 ppm
Feed Size Flotation Recovery (%)
KMnO4 and lime, electroflotation resulted in significantly
Mineral (µm) Condtions EF DAF MF Reference
higher pyrite recovery as compared with dispersed-air flota-
Pyrite 25–45 15 ppm KEX; 55 82 Kydros et al.
pH 8; 10 ppm 1994 tion in the pH range 5.0–9.0 as shown in Figure 10b. The
CuSO4; depressing action of permanganate ion on pyrite flotation was
10 ppm lime claimed to be because of the formation of some surface
25–45 15 ppm KEX, 80 45 Kydros et al.
100 ppm 1994 products facilitated by the presence of permanganate ions,
KMnO4, not because of xanthate oxidation (Rinelli et al. 1980).
10 ppm lime; Oxidation of pyrite due to the presence of oxygen bubbles is
pH 8
−45 20 ppm KEX; 63 Matis et al. not expected to be significant in the presence of KMnO4.
50 ppm 1993
2+
Cu ; pH 8
Chalco- −20 140 mg/l KEX; 90 33 Bkaskar and 4.1.2 Control of Bubble Sizes and Bubble Flux
pyrite pH 8.0; Khangaonkar,
4% Na2SO4, O2 1982 Bubble size and bubble flux can be controlled by adjusting
gas pulp pH, current density, pulp composition (e.g., type and
−20 0 KEX; pH 11; > 75 Trahar 1981 concentration of reagents, etc.), electrode material, and elec-
PPG
10–22 wt% trode surface geometry (Bhaskar Raju and Khangaonkar 1982;
solids Ren et al. 2014b). High concentration of ultrafine bubbles
Sphalerite −25 pH 4; O2 gas; 70 50 Llerena et al. provides for better bubble distribution within a flotation cell
192 mg/L 1996
SIX; 123 mg/L which is not possible to achieve in mechanically agitated
CuSO4 flotation cells. It should be noted that given that mechanical
−25 pH 4; H2 gas; 95 85 Llerena et al. agitation is not used in electroflotation, mechanical entrain-
192 mg/L 1996
SIX; 123 mg/L ment of fine particles can be neglected.
CuSO4
−10 22 g/t KEX; 57 Trahar 1976
20 wt% solids; 4.1.3 Use of Electrolytic Hydrogen and Oxygen Bubbles
pH 11; 16 min
flotation
Hydrogen and oxygen bubbles, occurring during electrolysis, are
Quartz 3–15 4.46 × 10−5 M > 82 Sarkar et al. more active than air bubbles generated during dispersed-air
CTAB; pH 10 2011 flotation because oxygen and hydrogen can be adsorbed on
MIBC; 0.2 M
Na2SO4
mineral surfaces, causing various electrochemical reactions par-
10–30 2.5 mg/L 94 Ketkar et al. ticularly on surfaces of sulphide minerals (Venkatachalam 1992).
sodium oleate; 1991 As a result, the presence of electrolytically generated oxygen and
H2 gas; 0.5%
solids; pH 9.0 hydrogen bubbles might be beneficial to recover specific
4–10 2.5 mg/L 70 Ketkar et al. minerals.
sodium oleate; 1991 Another advantage of electroflotation is that hydrogen and
H2 gas; 2.5%
solids; pH 9.0 oxygen bubbles can be isolated and used separately for differ-
−10 TMCS; pH 6; 70 Shi and ent flotation conditions. For instance, it was reported that the
3.5 min Fornasiero floatability of sulphide minerals such as chalcopyrite (Bhaskar
flotation 2009
Dolomite 10–45 20 mg/L fatty 90 74 Matis et al. Raju and Khangaonkar 1982, 1984b) or pyrite (Glembotsky
acid; pH 10 1993 et al. 1975; Kydros et al. 1994) in the presence of xanthate
15–45 100 ppm CMC; 70 Matis et al.
pH 9.5 1993
collector were higher with electrolytically generated oxygen
bubbles as compared with hydrogen bubbles. The reason is
KEX = potassium ethyl xanthate; CTAB = Cetyl trimethyl ammonium bromide;
CMC = Carboxymethyl cellulose; SIX = sodium isopropyl xanthate; that oxygen is required for the formation of dixantogen or
PPG = polypropylene glycol; TMCS = trimethylchlorosilane metal xanthate on surfaces of sulphide minerals, rendering
EF = Electroflotation; DAF = Dissolved-air flotation; MF = mechanical flotation these minerals hydrophobic. It was reported that the rate of
adsorption of xanthate on pyrite and chalcopyrite increased
with increasing the oxygen concentration, even at relatively
potassium ethyl xanthate and 10 ppm copper sulphate, flota- high oxygen concentrations (Kuopanportti et al. 1997). Thus,
tion recovery obtained using electrolytic bubbles was lower the presence of dissolved oxygen facilitates the adsorption of
than that obtained by sparging bubbles in similar conditions collectors and increases sulphide mineral recovery, although
as seen in Figure 10a. The reason for that might be the prolonged oxidation could depress sulphide mineral flotation.
decomposition of copper (I) xanthate to copper (II) oxide at The pulp potential was also found to be strongly dependent
high oxidation potential during electroflotation, making pyrite on the dissolved oxygen concentration for both pyrite and
particles less hydrophobic (Woods et al. 1990). Depression of chalcopyrite. Generally, higher dissolved oxygen concentra-
pyrite flotation was observed at pH 3–6 range while activation tion led to more positive pulp potential (Kuopanportti et al.
of pyrite was found at pH > 6 for both electroflotation and 1997). It was suggested that the pulp potential may be the
dispersed-air flotation. The sinusoidal shape of the flotation driving force in xanthate adsorption in the flotation of sul-
recovery of pyrite in various pH conditions shown in phide minerals such as pyrite and chalcopyrite.
MINERAL PROCESSING AND EXTRACTIVE METALLURGY REVIEW 9

100 100
a b
Electroflotation
Electroflotation
80 80
Recovery (%)

Recovery (%)
60 Dispersed-air flotation 60

40 40 Dispersed-air flotation

20 20

0 0
5 6 7 8 9 10 11 12 5 6 7 8 9 10 11 12
pH pH

Figure 9. Comparison between electroflotation (open circles) and dispersed-air flotation (filled circles) recoveries of (a) dolomite; (b) magnesite at various pulp pH;
both at particle size fraction of −45 + 10 µm and in the presence of 20 mg/L fatty acid collector (Matis et al. 1993).

100 100
a b Electroflotation
Electroflotation
Dispersed-air flotation Dispersed-air flotation

80 80
Recovery (%)
Recovery (%)

60 60

40 40

20 20

0 0
2 4 6 8 10 12 14 2 4 6 8 10 12 14
pH pH

Figure 10. Comparison between electroflotation and dispersed-air flotation recoveries of fine pyrite (a) in the presence of 15 ppm KEX and 10 ppm CuSO4; and (b) in
the presence of 15 ppm KEX, 100 ppm KMnO4 and lime at different pulp pH (Kydros et al. 1994).

In regards to oxide minerals such as cassiterite (Hogan Energy per ton of floatable particles
et al. 1979) and quartz (Lumanauw 2000), hydrogen bub- Cell voltage ðV Þ  Current ðAÞ  timeðhÞ
bles led to higher flotation recoveries as compared with ¼ (16)
weight of particles floated ðtÞ
oxygen gas bubbles. In the case of electroflotation of cassi-
terite with electrolytic hydrogen bubbles, increased flotation Based on Equation (16), and at a current of 5 A, the cell voltage
recovery and grade was observed due to surface reduction of 9.8 V, flotation time of 4 min, and using electrolytic hydrogen
of cassiterite by electrolytic hydrogen, making cassiterite at pH 4, they estimated the energy requirement to be equal to 65
surfaces more hydrophobic (Hogan et al. 1979). kWh/t for electroflotation of ZnS (Llerena et al. 1996). A sig-
nificantly lower estimated energy consumption of 3 kWh/t was
reported by the same authors for conventional flotation.
The generation of hydrogen ions at the cathode changes the
4.2 Disadvantages of Electroflotation
overall pH of the pulp around the vicinity of the electrodes
The main disadvantage of electroflotation is that it requires (Glembotsky et al. 1975; Llerena et al. 1996). Although this
the use of electrical energy for bubble generation which localized change in pH may not be significant if constant mixing
increases production costs (Montes-Atenas et al. 2010). is maintained, mineral surface hydrophobicity, dissociation rate
The energy consumption during electroflotation depends of reagents, and ionic composition of pulp may all be influenced
on a range of factors such as the nature of electrodes by small changes in pulp pH. The tight control of pulp pH is
used and the distance between them, electrolyte conductiv- needed because uncontrolled fluctuations in pulp pH have detri-
ity, pulp density, floatability of the particles, pH, and also mental effects on flotation performance. Furthermore, at lower
operational factors to a large extent. Nevertheless, Llerena pulp pH, metals may leach out of the minerals (Bhaskar Raju and
et al. expressed the energy consumption during electroflo- Khangaonkar 1982; Ren et al. 2014b) and stability of the flotation
tation as (Llerena et al. 1996): reagents may deteriorate, substantially reducing mineral
10 B. TADESSE ET AL.

recovery (Llerena et al. 1996). It is also worth mentioning that, in The average bubble loading parameter can be determined
cases where the electrolysis electrolyte is not separated from the using the following expression (Koh and Schwarz 2006):
pulp, the presence of inorganic salts in flotation pulp might have
a depressing effect on the flotation of some particles. 4dp 2 Np;a
β¼ (20)
Sulphide minerals are prone to surface oxidation when db 2 φNbT
exposed to oxygen, and their flotation is greatly dependent where dp and db are the particle and bubble size, respectively;
on the surface chemistry of the mineral particles in the pulp. Np;a is the number of particles attached to bubbles, and φ is
This is particularly of interest in electroflotation because oxy- the fractional coverage of bubble surface. In practice, the
gen gas is generated at the anode. Electrochemical reactions whole bubble surface may not be completely covered by
involving oxygen may be expected on the surface of mineral particles, and thus φ can be assumed to be 0.5. However,
particles especially in a set up where the electrodes are within Koh and Schwartz (2008) compared the experimental and
the flotation chamber. For instance, it is reported that sul- calculated recovery and found that the value of φ of 0.2
phide minerals undergo physicochemical changes in an elec- matched experimental data better than a value of 0.5.
troflotation cell (Glembotsky et al. 1975). Chalcopyrite Equation (19) can be used to calculate the number of
undergoes electrochemical oxidation in alkaline conditions particles attached to bubbles when the
and forms thin layer of Cu(OH)2, Fe(OH)3, CuO, and Fe2O3 bubble–particle attachment rate constant is known.
on its surface (Yin et al. 2000), particularly in cases where Therefore, in this section, a brief description of calculation
bubbles are generated in the pulp. of bubble–particle attachment rate constants is given.
CuFeS2 þ 5OH  ! CuðOH Þ2 þ FeðOH Þ3 þ 2S þ 5e (17) Additionally, the available models for prediction of the max-
imum floatable particle size in electroflotation are also
2CuFeS2 þ 10OH  ! 2CuO þ Fe2 O3 þ 5H2 O þ 4S presented.
þ 10e (18)
These surface changes may reduce flotation recovery signifi-
cantly. Additionally, when common sulphide collectors such 5.1. Bubble–Particle Attachment Rate Constant
as xanthates are used, xanthate might decompose to a certain The bubble–particle attachment rate constant can be pre-
degree, reducing the ability of xanthate to make sulphide dicted using the theory by Koh and Schwarz (2006):
minerals sufficiently hydrophobic. Due to chalcopyrite surface
oxidation to Fe oxides, and possible xanthate decomposition k1 ¼ Z 1 P (21)
during electroflotation, collectors such as sodium oleate which where Z1 is the bubble–particle collision frequency, and P is
are normally used in Fe oxide flotation (Joseph-Soly et al. the probability of collection of particles by bubbles.
2015) were found to be effective in the electroflotation of Given that there is no mechanical agitation during electro-
ultrafine chalcopyrite (Makuei et al. 2018). Thus, it can be flotation, Z1 can be expressed by (Nguyen and Schulze 2004):
concluded that electroflotation may be more suitable in the  
flotation of oxide minerals such as hematite, magnetite, and db þ dp 2
Z1 ¼ π Vrel (22)
other carbonate and silicate minerals where oxidation is not a 2
primary concern, as compare with sulphide minerals such where Vrel is the relative velocity between the bubble (vb ), and
chalcopyrite and pyrite. the particle (vp ).
The probability of collection can be expressed as (Yoon
2000):
5. Electroflotation Models P ¼ P c P a ð1  P d Þ (23)
The first-order flotation model can be used to predict the rate of where Pc ; Pa ; andPd are the probability of collision, attach-
recovery of mineral particles in electroflotation cells. Recovery of ment, and detachment, respectively.
particles by flotation depends on the number of particles col- For fine particles, Pd can be negligibly small due to the low
lected in the concentrate which in turn is a function of the inertia of fine particles, and thus Equation (23) may be sim-
number of free particles in the feed, the number of particles plified as:
attached to bubbles, and bubble–particle detachment rate. The
bubble–particle detachment rate can be neglected in electroflo- P ¼ Pc Pa (24)
tation processes because there is no agitation and fine particles
have very small mass (Yoon and Luttrell 1989). The number of 5.1.1. Bubble–Particle Collision
particles attached to bubbles (Np;a Þ in an electroflotation cell can The relationship between the probability of collision, and bubble
be expressed as (Koh and Schwarz 2006): and particle sizes has been a subject of extensive research, and
t several mathematical models have been developed (Flint and
Np;a ðt Þ ¼ ò k1 Np;f NbT ð1  βÞ (19) Howarth 1971; Gaudin et al. 1942; Sutherland 1948; Reay and
0
Ratcliff 1973; Yoon 1993). For very small bubbles (< 100 μm)
where Np;f is the number of free particles, k1 (m3/s) is the such as those encountered in electroflotation, Stokes flow might
bubble–particle attachment rate constant and NbT is the total occur around air bubbles. In the case of Stokes flow, the prob-
number of bubbles. β is the average bubble loading parameter. ability of collision is directly proportional to the square of the
MINERAL PROCESSING AND EXTRACTIVE METALLURGY REVIEW 11

particle size and inversely proportional to the square of the particle attachment is only thermodynamically feasible if θ
bubble diameter (Yoon 2000). > 0. The contact angle is used to describe the hydrophobicity
 2 of mineral surfaces.
dp Bubble–particle aggregates occur when hydrophobic parti-
Pc ¼ 1:5 (25)
db cles collide with bubbles, followed by the drainage and rup-
Intermediate flow around air bubbles can also occur during ture of the liquid film between the bubble and the particle,
electroflotation due to the mixing action of electrolytic bub- and the expansion and relaxation of the gas–liquid–solid
bles. The probability of collision can be derived using the contact line (Ralston et al. 1999). The time scale of the
stream function (Weber and Paddock 1983; Ralston et al. drainage of the liquid film is defined as the induction time.
1999): Given that the drainage of the liquid film controls the bubble–
  2 particle attachment mechanism, the induction time is
3 4Re0:72 dp required to determine the probability of attachment. For
Pc ¼ þ b
(26)
2 15 db example, Yoon and Luttrell (1989) used the stream function
and the induction time to derive the following expression for
Reb is the Reynolds number of bubbles.
the probability of attachment:
The decrease in particle size and the increase in bubble size
result in lower Pc, suggesting that fine particles are less likely "  ! #
1  45 þ 8Re0:72
b vb ti
to be captured by large bubbles. By contrast, fine bubbles Pa ¼ sin 2tan exp
2   (29)
enhance the probability of collision with fine particles which 15db 1 þ db =dp
leads to increasing the probability of collection.
where ti is the induction time and vb is the bubble rise
velocity.
5.1.2. Bubble–Particle Attachment Ren et al. (2014a) investigated bubble–particle attach-
Bubble–particle attachment is the central process in froth ment mechanism between fine cassiterite particles and elec-
flotation (Ralston et al. 1999; Verrelli and Albijanic 2015). trolytic hydrogen bubbles and found that the presence of
Flotation thermodynamics is used to predict the possibility small bubbles may cause flocculation of fine particles
of the attachment of a particle to a bubble. In electroflota- through a bubble-bridging mechanism by capillary action
tion, where solid particles and bubbles are suspended in (Figure 11a). Thus, the apparent particle size increases,
liquid medium, the total surface free energy is equal to the leading to higher probability of bubble–particle collision.
balance between the solid, liquid, and vapor interfacial In addition to this, fine particles might be coated with
tensions and interfacial areas. The relationship between electrolytic microbubbles thereby modifying surface proper-
the Gibbs free energy and the three surface forces is given ties of the particles (Figure 11b). It was suggested that these
by (Laskowski 1989): bubbles act as flotation carriers, and lead to improved
ΔG ¼ γsv  γsl  γlv (27) flotation efficiencies (Ahmadi et al., 2014; Tussupbayev
et al. 2016). For instance, the effect of microbubbles on
where is γsv the solid–vapor interfacial tension, γsl is the the selective flotation of ultrafine chalcopyrite (80% passing
liquid–solid interfacial tension, γlv is the liquid–vapor inter- 12 µm, 50% passing 4 µm) using laboratory pneumomecha-
facial tension, respectively. The Gibbs free energy change for nical and column flotation cells was investigated by
attachment must be negative for a bubble–particle attachment Tusupbajev et al. (2016). In comparison with the absence
to be feasible thermodynamically. Equation (27) can be con- of microbubbles, the introduction of 16 L of microbubbles
verted, with the introduction of Young’s equation, to the well- per 1 kg of chalcopyrite in a pneumomechanical type cell
known thermodynamic criterion for flotation: resulted in an increase in the copper recovery from 60% to
72%. Similarly, it was observed that the presence of micro-
ΔG ¼ γlv ðcosθ  1Þ (28)
bubbles in flotation column increases the copper recovery
where θ denotes contact angle between the bubble and parti- by 15.8% compared with conventional column flotation
cle surface. From Equation (28) it can be seen that bubble– (Tusupbajev et al., 2016).

Figure 11. Schematic representation of possible bubble–particle attachment mechanism in electroflotation of fine particles: (a) flocculation of particles by bubble
bridging; (b) particles coated with fine bubbles; and (c) two-stage bubble–particle attachment (Zhou et al. 1997).
12 B. TADESSE ET AL.

Fine particles coated with ultrafine bubbles may attach to a more active than those generated by dispersed-air flotation
large bubble, resulting in the formation of a bubble–bubble– methods. The particle surface modifications resulting from inter-
particle aggregate (Dziensiewicz and Pryor 1950) as shown in action with electrolytic gas bubbles were reported to result in
Figure 11c. The attachment of large bubbles to a mineral particle better floatability of pyrite and chalcopyrite even in the absence
frosted with tiny bubbles, typically in the range of about 1–10 µm, of collectors (Glembotsky et al. 1975). Another important feature
is highly probable because tiny bubbles forming on the particle of electroflotation is its ease of operation. Electroflotation can be
surface facilitate the attachment of larger bubbles (Klassen 1960). operated in combination with electrocoagulation particularly in
A range of bubble sizes with diameter from < 10 µm to > 100 µm fine and ultrafine particles flotation (Chen 2004). In such con-
are known to form in electroflotation and thus this mechanism figuration, the electrocoagulation process mainly plays the role
may be favored for a number of reasons. For instance, particle of destabilizing and aggregating the fine particles, while the
surface frosted by bubbles is expected to be less hydrated as electroflotation part is used to float the resulting flocs. A com-
compared with the surface of mineral particles because bubbles bined electrocoagulation-electroflotation system has been found
are more hydrophobic than solid particles. Thus, the coalescence to be effective for large-scale applications (Chen 2004).
of a bubble to a particle covered by bubbles is more favored than Although electroflotation can lead to improved selectiv-
the attachment of a bubble to a solid particle. ity and recovery, small bubbles cause high water recovery,
and the entrainment of gangue minerals, although mini-
mized, cannot be avoided altogether (Trahar and Warren
5.2. Limits of Particle Size in Electroflotation
1976). In acidic pH, metallic ions may leach out in to
The maximum floatable particle size was derived using the solution during electroflotation and may form species that
balance of the capillary force, the pressure force, the buoyancy deactivate the surface of minerals, consequently depressing
force, and the gravitational force (Nguyen 2003): flotation recovery (Llerena et al. 1996). Thus, electroflota-
 1 tion may not be as effective in low pH range. Although
  3σ ð1  cosθÞ 2 large-scale electroflotation cells have been designed for
dp max;d ¼ (30)
Δρðg þ bm Þ waste water treatment, similar scale cells for mineral parti-
cle recovery purposes are scarce. Furthermore, the greater
where dp max is the maximum diameter of floatable particles,
energy requirement during water electrolysis for bubble
Δρ ¼ ρp  ρl , ρp and ρl are the particle density (kg/m3) and
generation poses constraints to its widespread use for
the liquid density (kg/m3), respectively, θ is the particle con- mineral beneficiation on an industrial scale (Bhaskar Raju
tact angle, σ is the liquid surface tension (N/m), g is the and Khangaonkar 1984b). However, the following factors
acceleration due to gravity (m/s2) and bm is the machine were identified as compensation for the extra expenditures
acceleration (m/s2). incurred by electroflotation (Glembotsky et al. 1975):
Given that machine acceleration is negligible in electroflo-
tation, Equation (30) can be simplified as (Sarkar 2012): (1) Lowering or elimination of collector consumption
 1 (e.g. 98% recovery of pyrite during electroflotation
  3σ ð1  cosθÞ 2
dp max;d ¼ (31) with oxygen bubbles were obtained without any
gΔρ reagent addition (Glembotsky et al. 1975); high chal-
Therefore, if θ, Δρ and σ are known the maximum particle copyrite recovery was achieved in the absence of
size can be calculated. collector during electroflotation (Bhaskar Raju and
Another model to predict the maximum floatable particle Khangaonkar 1984a)).
size is derived for the case when the mass of the particles (2) Improved flotation rate
attached to the bubble prevents the ability of the bubble– (3) Improved grade of target mineral in the concentrate
particle aggregate to rise into the froth phase (Wark 1933). (4) Applications in non-conventional flotation separa-
More precisely, the model is based on the buoyancy criteria: tions (e.g. recovery of fine salt-type minerals (Matis
!1 et al. 1993); removal of toxic heavy metals from
  ρpulp  ρg 3 wastes (Il’in et al. 2014); and separation of microbes
dp max;b ¼ db (32) from wastewater (Neto et al. 2014)).
ρp  ρpulp

where ρg; ρpulp ; ρp are the density of gas, pulp and particle,
respectively. Thus, Equation (32) shows that the maximum
size of floatable particle can be estimated when
7. Conclusions
ρg; ρpulp ; ρp anddb are known. The challenges in the froth flotation of fine and ultrafine
particles are well recognized. Several experimental and theo-
retical studies have been published aiming at improving the
6. Future Outlook of Electroflotation in Fine Particle
recovery of this particle size range. The main cause for
Recovery
reduced recovery of fine and ultrafine particles in conven-
The advantages of electroflotation over dispersed-air flotation lie tional flotation methods is the poor collision efficiency
not only in the smaller bubble sizes generated by electrolysis but between the fine particles and large bubbles. Methods invol-
also in the ease of control bubble sizes by varying operating ving small bubbles are considered to be more suitable for the
variables. Gas bubbles generated in electroflotation are also recovery of ultrafine particles. Electroflotation has been
MINERAL PROCESSING AND EXTRACTIVE METALLURGY REVIEW 13

proved to be a promising method in the beneficiation of fine Chen, X., Chen, G., and Yue, P. L., 2002, “A novel electrode system for
and ultrafine mineral particles because of its ability to produce electro-flotation of wastewaters.” Environmental Science and
Technology, 36. pp. 778–783
gas bubbles of < 100 µm in size. Compared with dispersed-air
Dai, Z., Dukhin, S., Fornasiero, D., and Ralston, J., 1998, “The inertial
flotation, electroflotation has been found to be more effective hydrodynamic interaction of particles and rising bubbles with mobile
in the recovery of carbonates such as dolomite and magnesite. surfaces.” Journal of Colloid and Interface Science, 197. pp. 275–292
Recoveries in excess of 90% have been reported for the elec- Dai, Z., Fornasiero, D., and Ralston, J., 2000, “Particle-bubble collision
troflotation of ultrafine silica particles using electrolytic models – a review.” Advances in Colloid and Interface Science, 85. pp.
231–256
hydrogen bubbles. A two-stage bubble–bubble–particle
Dobby, G. S., and Finch, J. A., 1987, “Particle size dependence in flota-
attachment is proposed as the main attachment mechanism tion derived from a fundamental model of the capture process.”
in electroflotation. Despite its advantages in the ease of bubble International Journal of Mineral Processing, 21. pp. 241–260
size control, electroflotation requires additional energy for gas Dziensiewicz, J., and Pryor, E. J., 1950, “An investigation into the action
bubble generation. This additional energy cost needs to be of air in froth flotation.” Transactions of the Institutions of Mining and
Metallurgy, 59. pp. 455–491
adequately compensated by the improved recovery in order to
Elmore, F. E., 1905, “A process for separating certain constituents of
realize the application of electroflotation on an industrial subdivided ores and like substances and apparatus therefore.” British
scale. Further research on exploring equipment design Patent, 13. pp. 578
requirements is essential. Flint, L. R., and Howarth, W. J., 1971, “The collision efficiency of small
particles with spherical air bubbles.” Chemical Engineering Science, 26.
pp. 1155–1168
Disclosure statement Fuerstenau, D. W., 1980, Fine particle flotation, P. Somasundaran, Ed.,
Fine Particles Processing, New York: AIME
The authors report no conflicts of interest. The authors alone are Fuerstenau, D. W., and Mishra, B. K., 1981, On the mechanism of pyrite
responsible for the content and writing of the article. notation with xanthate as collectors, M. H. Jones, Editor, Complex
Sulfides, London: IMM, 271–278
Fuerstenau, M. C., Kuhn, M. C., and Elgillani, D. A., 1968, “The role of
Funding dixanthogen in xanthate flotation of pyrite.” Transactions of AIME,
241. pp. 148–156
The financial assistance of the Department of Mining Engineering and
Gaudin, A. M., Groh, J. O., and Henderson, H. B., 1931, “Effect of
Metallurgical Engineering, Western Australian School of Mines, Curtin
particle size on flotation.” AIME Technical Publications, 414. pp. 3–23
University is gratefully acknowledged.
Gaudin, A. M., Schuhmann, R., and Schlechten, A. W., 1942, “Flotation
kinetics. II. the effect of size on the behaviour of galena particles.”
Journal of Physical Chemistry, 46. pp. 902–910
References Glas, J. P., and Westwater, J. W., 1964, “Measurements of the growth of
electrolytic bubbles.” International Journal of Heat and Mass Transfer,
Ahmadi, R., Khodadadi, D. A., Abdollahy, M., and Fan, M., 2014, “Nano-
7. pp. 1427–1430
microbubble flotation of fine and ultrafine chalcopyrite particles.”
Glembotsky, V. A., Mamakov, A. A., Romanov, A. M., and Nenno, V. E.,
International Journal of Mining Science and Technology, 24. pp. 559–
1975, Selective separation of fine mineral slimes using method of
566
electric flotation. in: Proceedings of 11th International Mineral
Al Shakarji, R., He, Y., and Gregory, S., 2011, “The sizing of oxygen
Processing Congress, Cagliari, pp. 561–582.
bubbles in copper electrowinning.” Hydrometallurgy, 109. pp. 168–174
Han, M. Y., Park, Y. H., and Yu, T. J., 2002, “Development of a new
Alexandrova, L., Nedialkova, T., and Nishkov, I., 1994, “Electroflotation
method of measuring bubble size.” Water Science and Technology, 2.
of metal ions in waste water.” International Journal of Mineral
pp. 77–83
Processing, 41. pp. 285–294
Hogan, P. A., Kuhn, T., and Turner, J. F., 1979, “Electroflotation studies
Ben Mansour, L., Chalbi, S., and Kesentini, I., 2007, “Experimental study
based on cassiterite ores.” Transactions of the Institute of Mining and
of hydrodynamic and bubble size distributions in electroflotation
Metallurgy, 88. pp. C83–C87
process.” Indian Journal Chemical Technology, 14. pp. 253–257
Il’in, V. I., Perfil’eva, A. V., and Kolesnikov, V. A., 2014, “Effect of
Bhaskar Raju, G., and Khangaonkar, P. R., 1982, “Electro-flotation of
magnetic field on the electroflotation extraction of sparingly water-
chalcopyrite fines.” International Journal of Mineral Processing, 9. pp.
soluble heavy and non-ferrous metal compounds from wastewater.”
133–143
Russian Journal of General Chemistry, 84. pp. 2315–2319
Bhaskar Raju, G., and Khangaonkar, P. R., 1984a, “Electroflotation-a
Janssen, L. J. J., and Hoogland, J. G., 1973, “The effect of electrolytically
critical review.” Transactions of the Indian Institute of Metals, 37.
evolved gas bubbles on the thickness of the diffusion layer-ii.”
pp. 59–66
Electrochimica Acta, 18. pp. 543–550
Bhaskar Raju, G., and Khangaonkar, P. R., 1984b, “Electroflotation of
Jiménez, C., Talavera, B., Sa’ez, C., Can’izares, P., and Rodrigo, M. A.,
chalcopyrite fines with sodium diethyldithiocarbamate as collector.”
2010, “Study of the production of hydrogen bubbles at low current
International Journal of Mineral Processing, 13. pp. 211–221
densities for electroflotation processes.” Journal of Chemical
Brandon, N. P., and Kelsall, G. H., 1985, “Growth kinetics of bubbles
Technology and Biotechnology, 85. pp. 1368–1373
electrogenerated at microelectrodes.” Journal of Applied
Jones, S. F., Evans, G. M., and Galvin, K. P., 1999a, “Bubble nucleation
Electrochemistry, 15. pp. 475–484
from gas cavities -a review.” Advances in Colloid and Interface Science,
Brandon, N. P., Kelsall, G. H., Levine, S., and Smith, A. L., 1985,
80. pp. 27–50
“Interfacial electrical properties of electrogenerated bubbles.” Journal
of Applied Electrochemistry, 15. pp. 485–493 Jones, S. F., Evans, G. M., and Galvin, K. P., 1999b, “The cycle of bubble
Burns, S. E., Yiacoumi, S., and Tsouris, C., 1997, “Microbubble genera- production from a gas cavity in a supersaturated solution.” Advances
tion for environmental and industrial separations.” Separation and in Colloid and Interface Science, 80. pp. 51–84
Purification Technology, 11. pp. 221–232 Joseph-Soly, S., Quast, K., and Connor, J. N., 2015, “Effects of Eh and pH
Chen, G., 2004, “Electrochemical technologies in wastewater treatment.” on the oleate flotation of iron oxides.” Minerals Engineering, 83. pp.
Separation and Purification Technology, 38. pp. 11–41 97–104
Chen, X., and Chen, G., 2010, Electroflotation, C. Comninellis and G. Ketkar, D. R., Mallikarjunan, R., and Venkatachalam, S., 1991,
Chen, Eds, Electrochemistry for the Environment, New York: Springer, “Electroflotation of quartz fines.” International Journal of Mineral
pp.263–277 Processing, 31. pp. 127–138
14 B. TADESSE ET AL.

Klassen, V. I., 1960, Theoretical basis of flotation by gas precipitation. Reay, D., and Ratcliff, G. A., 1973, “Removal of fine particles from water
The Proceedings of Vth IMPC, London, pp. 309–322. by dispersed air flotation- effects of bubble size and particle size on
Koh, P. T. L., and Schwarz, M. P., 2006, “CFD modelling of bubble- collection efficiency.” Canadian Journal of Chemical Engineering, 53.
particle attachments in flotation cells.” Minerals Engineering, 19. pp. pp. 178–185
619–626 Ren, L., Zhang, Y., Qin, W., Bao, S., Wang, P., and Yang, C., 2014b,
Koh, P. T. L., and Schwarz, M. P., 2008, “Modelling attachment rates of “Investigation of condition-induced bubble size and distribution in
multi-sized bubbles with particles in a flotation cell.” Minerals electroflotation using a high-speed camera.” International Journal of
Engineering, 21. pp. 989–993 Mining Science and Technology, 24. pp. 7–12
Kubritskaya, T. D., Drako, I. V., Sorokina, V. N., and Drondina, R. V., Ren, L.-Y., Zhang, Y.-M., Qin, W.-Q., Bao, S.-X., and Wang, J., 2014a,
2000, “Use of electrochemical methods to purify the waste water from “Collision and attachment behavior between fine cassiterite particles
the production of concentrates in the food industry.” Surface and H2 bubbles.” Transactions of the Nonferrous Metals Society of
Engineering and Applied Electrochemistry, 6. pp. 62–68 China, 24. pp. 520−527
Kuopanportti, H., Suorsa, T., and Pollanen, E., 1997, “Effects of oxygen Rinelli, G., Marabini, A. M., and Alesse, V., 1980, “Depression action of
on kinetics of conditioning in sulphide ore flotation.” Minerals permanganate on pyrite and galena flotation” in “Complex Sulphide
Engineering, 10. pp. 1193–1205 Ores-Conference”, Rome, pp. 199–206.
Kydros, K. A., Gallios, G. P., and Matis, K. A., 1994, “Electrolytic flota- Sarkar, M. S. K. A., 2012, PhD Thesis on Electroflotation: Its application
tion of pyrite.” Journal of Chemical Technology and Biotechnology, 59. to water treatment and mineral processing, Newcastle, University of
pp. 223–232 Newcastle.
Kyzas, G. Z., and Matis, K. A., 2016, “Electroflotation process: a review.” Sarkar, M. S. K. A., Donne, S. W., and Evans, G. M., 2011, “Utilization of
Journal of Molecular Liquids, 220. pp. 657–664 hydrogen in electroflotation of silica.” Advanced Powder Technology,
Landolt, D., Acosta, R., Muller, R. H., and Tobais, C. W., 1970, “An 22. pp. 482–492
optical study of cathodic hydrogen evolution in high rate electrolysis.” Sarkar, M. S. K. A., Evans, G. M., and Donne, S. W., 2010, “Bubble size
Journal of Electrochemical Society, 117. pp. 839–845 measurement in electroflotation.” Minerals Engineering, 23. pp. 1058–1065
Laskowski, J. S., 1989, “Thermodynamic and kinetic flotation criteria.” Shi, Y., and Fornasiero, D., 2009, Effects of particle size and density, and
Mineral Processing and Extractive Metallurgy Review, 5. pp. 25–41 turbulence on flotation recovery, in: Proceedings of Chemeca, Perth, 1–9.
Llerena, C., Ho, J. C. K., and Piron, D. L., 1996, “Effects of pH Sides, P. J., and Tobias, C. W., 1985, “A close view of gas evolution from
on electroflotation of sphalerite.” Chemical Engineering the back side of a transparent electrode.” Journal of Electrochemical
Communications, 155. pp. 217–228 Society, 132. pp. 583–587
Lubetkin, S. D., 1995, “The fundamentals of bubble evolution.” Chemical Sivamohan, R., and Forssberg, E., 1985, “Recovery of heavy minerals
Society Reviews, 24. pp. 243–250 from slimes.” International Journal of Mineral Processing, 15. pp. 297–
Lumanauw, D., 2000, Masters Thesis on Hydrogen bubble characterisa- 314
tion in alkaline water electrolysis. Department of Metallurgy and Sutherland, K. L., 1948, “Physical chemistry of flotation - XI kinetics of
Material Science. Toronto, University of Toronto. the flotation process.” Journal of Physical Chemistry, 52. pp. 394–425
Makuei, F., Tadesse, B., Albijanic, B., and Browner, R., 2018, Trahar, W. J., 1976, “The selective flotation of galena from sphalerite
“Electroflotation of ultrafine chalcopyrite particles with sodium oleate with special reference to the effects of particle size.” International
collector.” Minerals Engineering, 120. pp. 44–46 Journal of Mineral Processing, 3. pp. 151–166
Matis, K. A., and Gallios, G. P., 1986, Dissolved-air and electrolytic Trahar, W. J., 1981, “A rational interpretation of the role of particle-size in
flotation, B. A. Wills and R. W. Barley, Eds., Mineral Processing at a flotation.” International Journal of Mineral Processing, 8. pp. 289–327
Crossroads, Dordrecht: Martinus Nijhoff Publishers Trahar, W. J., and Warren, L. J., 1976, “The flotability of very fine
Matis, K. A., Gallios, G. P., and Kydros, K. A., 1993, “Separation of fines particles -a review.” International Journal of Mineral Processing, 3.
by flotation techniques.” Separation Technology, 3. pp. 76–90 pp. 103–131
Matis, K. A., and Peleka, E. N., 2010, “Alternative flotation techniques for Tussupbayev, N. K., Rulyov, N. N., and Kravtchenco, O. V., 2016,
wastewater treatment: focus on electroflotation.” Separation Science “Microbubble augmented flotation of ultrafine chalcopyrite from quartz
and Technology, 45. pp. 2465–2474 mixtures.” Mineral Processing and Extractive Metallurgy, 125. pp. 5–9
Miettinen, T., Ralston, J., and Fornasiero, D., 2010, “The limits of fine Venkatachalam, S., 1992, “Electrogenerated gas bubbles in flotation.”
particle flotation.” Minerals Engineering, 23. pp. 420–437 Mineral Processing and Extractive Metallurgy Review, 8. pp. 47–55
Montes-Atenas, G., Garcia-Garcia, F. J., Mermillod-Blondin, R., and Verrelli, D. I., and Albijanic, B., 2015, “A comparison of methods for
Montes, S., 2010, “Effect of suspension chemistry onto voltage drop: measuring the induction time for bubble–particle attachment.”
application of electro-flotation.” Powder Technology, 204. pp. 1–10 Minerals Engineering, 80. pp. 8–13
Müller, L., Krenz, M., and Rübner, K., 1989, “On the relation between the Vogt, H., 1984, “The rate of gas evolution at electrodes-ii. an estimate of
transport of electrochemically evolved Cl2 And H2 into the electrolyte the efficiency of gas evolution on the basis of bubble growth data.”
bulk by convective diffusion and by gas bubbles.” Electrochimica Acta, Electrochimica Acta, 29. pp. 175–180
34. pp. 305–308 Wark, I. W., 1933, “The physical chemistry of flotation. I. the significance
Neto, R. G. C., Nascimento, J. G. S., Costa, M. C., Lopes, A. C., Neto, E. of contact angle in flotation.” Journal of Physical Chemistry, 37. pp.
F. A., Filho, C. R. M., and Santos, A. B., 2014, “Microalgae harvesting 623–644
and cell disruption: a preliminary evaluation of the technology elec- Weber, M. E., and Paddock, D., 1983, “Interceptional and gravitational
troflotation by alternating current.” Water Science and Technology, 70. collision efficiencies for single collectors at intermediate Reynolds
pp. 315–320 numbers.” Journal of Colloids and Interface Science, 94. pp. 328–335
Nguyen, A. V., 2003, “New method and equations for determining Woods, R., Young, C. A., and Yoon, R. H., 1990, “Ethyl xanthate
attachment tenacity and particle size limit in flotation.” International chemisorption isotherms and Eh-pH diagrams for the copper/water/
Journal of Mineral Processing, 68. pp. 167–182 xanthate and chalcocite/water/xanthate systems.” International
Nguyen, A. V., and Schulze, H. J., 2004, Colloidal Science of Flotation, Journal of Mineral Processing, 30. pp. 17–33
New York: Marcel Dekker, Inc Yin, Q., Vaughan, D. J., England, K. E. R., Kelsall, G. H., and Brandon,
Oliveira Da Mota, I., Castro, J. A., Góes Casqueira, R. A., and Oliveira N. P., 2000, “Surface Oxidation of Chalcopyrite (CuFeS2) in Alkaline
Junior, G., 2015, “Study of electroflotation method for treatment of Solutions.” Journal of Electrochemical Society, 147. pp. 2945–2951
wastewater from washing soil contaminated by heavy metals.” Journal Yoon, R.-H., 1993, “Microbubble Flotation.” Minerals Engineering, 6. pp.
of Materials Research and Technology, 4. pp. 109–113 619–630
Ralston, J., Fornasiero, D., and Hayes, R., 1999, “Bubble–particle attach- Yoon, R.-H., 2000, “The role of hydrodynamic and surface forces in
ment and detachment in flotation.” International Journal of Mineral bubble-particle interaction.” International Journal of Mineral
Processing, 56. pp. 133–164 Processing, 58. pp. 129–143
MINERAL PROCESSING AND EXTRACTIVE METALLURGY REVIEW 15

Yoon, R.-H., and Luttrell, G. H., 1989, “The effect of bubble size on fine Zouboulis, A. I., and Matis, K. A., 1989, “Electrolytic flotation of chro-
particle flotation.” Mineral Processing and Extractive Metallurgy mium from dilute solutions.” Environmental Technology Letters, 10.
Review, 5. pp. 101–122 pp. 601–612
Zhou, Z. A., Xu, Z., Finch, J. A., Hu, H., and Rao, S. R., 1997, “Role of Zouboulis, A. I., and Matis, K. A., 2012, “Cadmium ion removal by
hydrodynamic cavitation in fine particle flotation.” International electroflotation onto sewage sludge biomass.” International Journal
Journal of Mineral Processing, 51. pp. 139–149 of Environmental Waste Management, 9. pp. 245–256

You might also like