International Journal of Mineral Processing: Ali Vazirizadeh, Jocelyn Bouchard, Yun Chen

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

International Journal of Mineral Processing 157 (2016) 163–173

Contents lists available at ScienceDirect

International Journal of Mineral Processing

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

Effect of particles on bubble size distribution and gas hold-up in


column flotation
Ali Vazirizadeh a, Jocelyn Bouchard b,⁎, Yun Chen b
a
Mineral Processing, Hatch Ltd. Canada
b
Department of Mining, Metallurgical and Materials Engineering, LOOP (Laboratoire d'observation et d'optimisation des procédés), Université Laval, Quebec City, Canada

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

Article history: In flotation systems, it is generally assumed that the bubble size presents an inverse correlation with the gas hold-
Received 3 June 2016 up. However, this hypothesis is not systematically true. The focus of this work is to study the effect of solid particles
Received in revised form 21 October 2016 on the bubble size distribution and gas hold-up, as well as the correlation between the bubble size distribution and
Accepted 25 October 2016
gas hold-up in column flotation. Experiments were conducted in two and three-phase systems using a laboratory
Available online 28 October 2016
flotation column (5.6 cm internal diameter, total height of 650 cm), and mixtures of quartz (hydrophilic gangue)
Keywords:
and talc (naturally hydrophobic mineral), classified in four different size fractions. For the studied flotation system,
Bubble size distribution experimental results are compared with literature correlations and models to reveal that hydrophobic particles af-
Gas hold-up fect the gas hold-up through three different mechanisms modifying the Sauter mean diameter and rise velocity,
Column flotation namely (1) surface interactions, and the joint antagonistic effect of (2) bubble loading and (3) coalescence.
Solid particles © 2016 Elsevier B.V. All rights reserved.

1. Introduction the gas hold-up reduction in the presence of solids: (1) coalescence,
(2) a shift in the density and viscosity of the pulp, (3) a change in the ra-
In flotation systems, gas dispersion plays a critical role for particle dial gas hold-up and flow profiles, and (4) bubble wake effects. They
collection (recovery) and froth mass transport (selectivity). To evaluate concluded that changes in gas hold-up following the addition of solids
this effect at the industrial scale, the bubble surface area flux (Sb) is typ- were due to a combination of the two latter mechanisms, i.e. a change
ically estimated from the superficial gas rate (Jg) and bubble size distri- in radial gas hold-up and flow profiles, and bubble wake effects.
bution (BSD) measurements, where the complete BSD is compressed Gandhi et al. (1999) investigated the hydrodynamic behavior of
into a single value, i.e. the Sauter mean diameter (d32). However, it is slurry bubble columns at high solids concentrations, and found that
now criticized that (1) a given Sb value can be obtained from different the axial distribution of slurry concentration followed the classical sed-
combinations of Jg and d32 (Vinnett et al., 2012), and (2) a given d32 imentation–dispersion model. They also showed that the effect of the
value can be obtained from different BSDs (Maldonado et al., 2008b). gas velocity on axial solids distribution were minimal over the range
A better approach consists in adequately parameterizing the overall of gas velocities investigated.
BSD as presented by Vazirizadeh et al. (2015). To date, most of the laboratory-based research has been conducted
The gas hold-up is another important parameter used to character- on gas-water (two-phase) systems. Since one cannot conclude that re-
ize the hydrodynamic conditions of bubble column reactors (Luo et al., sults in two-phase systems necessarily apply for slurries, ‘surrogate’
1999). It is useful because it combines the influence of both the bubble solids such as talc, silica and coal have been used. Although such syn-
size and gas rate. It provides a holistic indication of the hydrodynamic thetic feed material are idealized/simplified compared to natural ores,
conditions because it is dependent on various factors such as frother they provide important insights about the nature of three-phase sys-
type and concentration, cell dimensions, operating temperature and tems in flotation (Kuan and Finch, 2010). Despite extensive research
pressure, as well as solid phase properties and concentration. on the effect of solids on gas-liquid systems, results are not conclusive
Banisi et al. (1995) observed that the presence of solid particles re- and sometimes apparently contradictory, for instance some of them
duces the gas hold-up in a column operated under conditions relevant showing a reduction in the gas hold-up (Banisi et al., 1995; Kuan and
to flotation. This effect increased in their experiments with solids con- Finch, 2010), and others an increase (Yang et al., 2007).
centration over the range 0–15% v/v. Hydrophilic (silica and calcite) Little is known about the physical mechanisms causing the observed
and hydrophobic (coal) particles produced similar reductions in the macroscopic effects. The majority of literature on the topic points to a
gas hold-up. They explored four possible mechanisms responsible for decrease in gas the hold-up with increasing solids concentration
(Banisi et al., 1995; Reese et al., 1996; Kuan and Finch, 2010). In other
⁎ Corresponding author. words, the presence of solids tends to increase the rise velocity of a bub-
E-mail address: jocelyn.bouchard@gmn.ulaval.ca (J. Bouchard). ble swarm.

http://dx.doi.org/10.1016/j.minpro.2016.10.005
0301-7516/© 2016 Elsevier B.V. All rights reserved.
164 A. Vazirizadeh et al. / International Journal of Mineral Processing 157 (2016) 163–173

Most of the literature focuses on the effect of solid particles on a sin- cells as described by Maldonado et al. (2008a). No wash water was
gle hydrodynamic parameter, either the gas hold-up (Banisi et al., 1995; used in the experiments. The air flow rate was measured and controlled
Kuan and Finch, 2010) or bubble size distribution (O'Connor et al., 1990; using a mass flow controller, whose readings were converted to volu-
Gandhi et al., 1999; Grau and Heiskanen, 2005). In this paper, the effect metric flow rates at reference conditions of gas flow rate, which were
of solid particles is studied not only on the bubble size distribution and then corrected for actual (test) temperature and pressure conditions as
gas hold-up, but also on the relationship between these two hydrody-   
namic characteristics jointly. Different mechanisms introduced from 1033:23 T þ 273:15
J g ¼ J ref
g ð2Þ
the literatures are discussed in light of experimental results. 1033:23 þ P 294:16

2. Experimental set-up where P is the measured absolute pressure (cm H2O) at the bottom of
column, T is the temperature (°C) and Jref g is the superficial gas rate
Experimental data were generated using a fully automated laborato- (cm/s) at 21 °C and 1033.23 cm H2O.
ry column flotation set-up depicted in Fig. 1. The column is made of The frit-and-sleeve sparger consisted of a porous stainless steel ring
5.6 cm internal diameter acrylic tubes, and exhibits a total height of concentrically installed within a cylindrical sleeve (Kracht et al., 2005).
650 cm. The porous ring was 4 cm long and 2.5 cm in diameter and had a nom-
A frit-and-sleeve sparger, detailed below, was located at the bottom inal pore diameter of 10 μm. The cylindrical gap (annulus) between the
of the column to provide steady air flow rates through local PID control- frit and the sleeve was 1 mm. This design allowed an external flow of
lers. Temperature and absolute pressure sensors were mounted at the water to circulate through the gap creating the necessary shear to re-
bottom of the column to compensate for standard condition flow rate duce the size of the bubbles generated in the device (Maldonado et al.,
measurements (referred to a given temperature and pressure). A con- 2015).
ductivity-based sensor (Gomez et al., 2003), installed in the middle of This kind of sparger thus provides an additional degree-of-freedom
the column (350 cm from the tailings port), allowed measuring the to control the bubble size with the water velocity, i.e. Jsl, defined such as:
gas hold-up using the conductivity of both the aerated and de-aerated
slurry. Maxwell's equation relates the conductivity to the concentration Q sl
J sl ¼ ð3Þ
of a dispersed non-conducting phase (i.e. bubbles) in a continuous liq- Agap
uid phase (pulp in this case) as:
where Qsl is the volumetric water flow rate in the sparger, and Agap is the
 cross sectional area of the gap. The water flow rate to the sparger is ma-
1−kd =kp
εg ¼  ð1Þ nipulated by the speed of a gear pump, using a PID loop implemented in
1 þ 0:5kd =kp
an industrial controller.
In order to continuously monitor the bubble size, the McGill bubble
where kd is the conductivity of the dispersed phase with air, and kp, the viewer (Hernandez-Aguilar et al., 2004) was adapted to the experimen-
conductivity of air-free dispersion (de-aerated slurry). tal column at 100 cm above feed port. The bubbles were collected via a
The column was operated under automatic control of froth-depth, sampling tube – 1.5 cm in diameter and oriented downward – and di-
gas rate, shearwater and feed flow rate. The froth zone level was mea- rected into a viewing chamber where they were exposed under lighting
sured and controlled using a sensor exhibiting eleven conductivity conditions to be photographed. The sampling tube diameter was ~ 7

Fig. 1. Schematic of the experimental set-up.


A. Vazirizadeh et al. / International Journal of Mineral Processing 157 (2016) 163–173 165

times larger than the largest observed bubble, which exceeds the gener- Table 1
ally accepted minimum opening (3 times the largest member of the Size classes for the talc feed material.

population) for particle sampling (Finch and Wills, 2015). Class 1 2 3 4


Parameters of the bubble size distribution – such as the Sauter mean Size (μm) +53/−75 +75/−106 +106/−150 +53/−150a
diameter, mean and variance at actual conditions (i.e. not corrected to a
The fourth class was prepared as a mix of equal amounts of the three other classes.
reference conditions) – were then determined by image analysis. With
this regard, the Circular Hough Transform (CHT)-based algorithm pre-
sented by Riquelme et al. (2013a, 2013b) was implemented for accurate
bubble detection (including large ones and clusters). Fig. 2a shows an
example of an original picture (24 pixels/mm) taken which contains For its constant efficiency over a rather long period of time, a
800 × 600 pixels. In Fig. 2b, the bubbles detected by the CHT-based polyglycol frother (F150) was selected. The superficial gas and feed ve-
algorithm are circled in blue and marked by a red cross in the middle. locities were set at 1 cm/s for all tests. Talc, as a natural hydrophobic
Three frames per second have been taken to ensure every picture pre- mineral consumes a fraction of F150, which is adsorbed on the particle
sented an entirely ‘new’ sample of bubbles, and avoid multiple counts surface (Kuan, 2009). To avoid the results being affected by this interac-
of the same bubble. Each test allowed sampling populations approxi- tion, samples started being taken after 1 h of column operation time,
mately composed of 10,000 bubbles, thus ensuring a representative thus ensuring constant frother efficiency over time.
distribution. All experiments were in closed circuit, i.e. the tailings and concen-
Quartz particles were used as a hydrophilic gangue trate streams were recirculated to the feed tank. Shearwater was
(+ 53 μm/− 150 μm) and talc, as a naturally hydrophobic mineral, drawn from the top of the feed tank through a strainer to avoid fine par-
ground and classified into four size classes presented in Table 1. The ticles to be entrained. The test procedure consisted in preparing the
pulp fed to the column contained 4% solids assaying 40% talc and 60% slurry feed with a given size fraction, and conducting the trials one
quartz. after the other with increasing frother concentration. The procedure
was repeated for every size fraction.

3. Results and discussion

(a) 3.1. Effect of solid particles on bubble size distribution

Fig. 3 provides the measured bubble size cumulative distribution


functions (CDF), and Fig. 4 shows the resulting bubble images as a func-
tion of solids addition in the presence of a polyglycol frother. The data at
0% solids (w/w) (i.e. the gas–water system) shows decreasing bubble
sizes (Fig. 3), and more uniform distributions (Fig. 4) with increasing
shearwater to the sparger.
The most significant impact of solids is to increase the bubble size.
This becomes apparent when introducing 4% (w/w) (40% talc and 60%
quartz) in the presence of 25 ppm F150 (i.e. above the critical coales-
cence concentration (CCC)): d32 increased from 0.72 mm to 1.35 mm
while introducing 10.1 cm/s shearwater (Jsw), and from 0.64 mm to
1.17 mm with 40.4 cm/s shearwater.
Kuan and Finch (2010) suggested that bubble coalescence in the
presence of talc was the explanation for this phenomenon. Bubble size
distribution measurements observed in the present investigation sup-
port this mechanism. For instance, Fig. 5 shows how the bubble size dis-
tribution evolves at 25 ppm F150 from (a) a narrow unimodal range in a
(b) water-only system to (b) a widespread one following the introduction
of talc particles (4% (w/w) slurry). This result is typical of coalescence:
the coexistence of fine and coarse bubbles.
At lower frother concentration (5 ppm F150), the bubble size distri-
bution shifts from unimodal to bimodal, which suggests a coalescence-
induced break-up mechanism (Tse et al., 2003). Quinn and Finch (2012)
also reported bimodal bubble size distributions in the absence of coales-
cence inhibiting agents, probably resulting from bubble-bubble interac-
tions. Fig. 6a depicts an example of a bimodal distribution typically
resulting from coalescence-induced break-up. At 5 ppm F150, the bub-
ble size distribution shifted following the introduction of solids
(4 wt%) to the feed tank, and the volume fraction of very fine bubbles,
i.e. b0.5 mm, then increased from 2.99 % to 28.63%. Bubble images
shown in Fig. 6b confirm that the population essentially consists in
two extreme classes, i.e. very fine and large bubbles.
Dippenaar (1982) reported that hydrophobic solids can promote co-
alescence through a bridging effect of hydrophobic particles between
Fig. 2. Example of bubble detection, (a) original image showing bubbles in a two-phase
bubbles. In addition, coalescence could occur because the frother con-
system (b) image with detected bubbles using the CHT-based algorithm, circled in blue centration decreases due to frother adsorption by talc (as a hydrophobic
and marked by a red cross in the middle. solid). With this regard, Kuan and Finch (2010) proposed the following
166 A. Vazirizadeh et al. / International Journal of Mineral Processing 157 (2016) 163–173

(a)

(b)

Fig. 3. Effect of the solids percent on the bubble size distribution (CDF) at different shearwater rate (a) 0 wt%, (b) 4 wt%.

hypothesis about the adsorption of F150 by talc particles based on the possibly exposing its hydrocarbon backbone with talc attached to bub-
structure of the frother molecule. Talc particles, via their uncharged sur- bles, thus inducing uptake by talc surfaces through hydrophobic
faces providing natural hydrophobicity, tend to attach to the bubbles. interactions.
On the other hand, the F150 molecule has two end OH groups (hydro- According to the above-mentioned hypothesis, talc surfaces exposed
philic) lying flat on the surface of the bubble to accommodate the two to bubble surfaces would play an important role on the adsorption of
hydrophilic groups (OH and H(C3H6O)). In this orientation, F150 is frother. Frother up-taken by particles is obviously not a desirable effect

(a) (b)

(c) (d)

1mm

Fig. 4. Examples of bubble images using 25 ppm F150, and for Jg = 1 cm/s. (a) Jsw = 10.1 cm/s, 0 wt% (b) Jsw = 40.4 cm/s, 0 wt%, (c) Jsw = 10.1 cm/s, 4 wt% (d) Jsw = 40.4 cm/s, 4 wt%.
A. Vazirizadeh et al. / International Journal of Mineral Processing 157 (2016) 163–173 167

(a)

(b)

Fig. 5. Effect of the solids percent on the bubble size histogram (25 ppm F150, Jg = 1 cm/s and Jsw = 40.4 cm/s). (a) 0 wt% (b) 4 wt%.

in flotation. However, it must be taken into account in cases where a concentration of frother in the solution, which remained over the CCC.
finer grind size is considered to improve the degree of liberation, as it Even for the first three tests, conducted using 5 ppm frother concentra-
will lead to a larger total surface of particles in the slurry. Similarly, in- tion, which is below the CCC of F150 (Finch et al., 2008), the particle size
creasing the solids fraction in a flotation circuit, for instance to increase effect could still not be observed, again most probably because the sur-
the production rate, could also lead to higher frother consumption. face area to volume ratio does not vary much with the particle size. A
To check the variation of frother up-take by increasing particle total different mineral system or experimental plan would be required to
available surface, a set of tests have been conducted. Four talc particle clarify the effect of frother up-take, using for instance the solids fraction
size-classes (53/75 μm, 75/106 μm, 106/150 μm and 53/150 μm) were as a factor, instead of the particle size distribution to modify the total
used, each one combined with the same quartz particle size class. The available hydrophobic mineral surface.
amount of both the hydrophobic and hydrophilic added solids remained
constant. For each particle class, three frother concentrations (5, 15 and 3.2. Effect of solid particles on gas hold-up
25 ppm) were applied in conjunction with three levels of shear water
rate (10.1, 20.2 and 40.4 cm/s). The shearwater rate provided an extra Different hypotheses have been postulated in the literature to ex-
degree-of-freedom for modifying the mean bubble size, particularly plain the influence of hydrophobic particles on gas hold-up. The effects
for tests conducted at a constant frother concentration. Table 2 summa- would be to:
rizes the experimental plan for each particle size class.
It is worth emphasizing that neither the effective frother concentra- • promote bubble coalescence leading to a reduction of the gas hold-up
tion nor the frother up-take was directly quantified in this work. The (Banisi et al., 1995; Kuan and Finch, 2010);
frother concentration was simply calculated from the amount of F150 • lead to bubble break-up, which in turn increases the gas hold-up
added to the volume of water. However, the proposed experimental de- (Yang et al., 2007);
sign allowed qualifying frother up-take indirectly, and therefore charac- • increase the apparent viscosity, and thus the gas hold-up (Finch and
terizing the effect. Dobby, 1990; Banisi et al., 1995);
Fig. 7 presents the effect of the talc particle size on the d32 at a con- • increase bubble wake entrainment effects which decrease the gas
stant gas rate of 1 cm/s. Coalescence leads to an increase of both the hold-up (Banisi et al., 1995);
d32 and standard deviation of bubble size distribution. As can be seen, • change the radial gas hold-up and flow profiles, thus leading to a re-
the particle size does not show a systematic effect on the d32, and there- duction of the global gas hold-up (Banisi et al., 1995);
fore, increasing the available surface of hydrophobic mineral surprising- • increase the weight of bubbles once attached, and therefore increase
ly did not always promote coalescence. the gas hold-up (Tsutsumi et al., 1991; Uribe-Salas et al., 2003).
It is worth mentioning with this regard that talc particles exhibit a
flattened shape, entailing that the surface area to volume ratio remains
approximately constant regardless of the particle size. This could ex- Each of these hypotheses is examined hereafter.
plain why finer mineral feed apparently does not seem leading to in-
creased frother up-take and bubble coalescence. 3.2.1. Bubble coalescence and break-up
Tests 4 to 9 in Table 2 were conducted using 15–25 ppm frother, Bubble coalescence occurs due to the bridging effect of particles
which is higher than the CCC of F150. Results could not reveal any signif- (Dippenaar, 1982) and/or adsorption of frother on the hydrophobic par-
icant effect of the particle size on coalescence, perhaps because of the ticle surface (Kuan and Finch, 2010). Faster rising larger bubbles thus
168 A. Vazirizadeh et al. / International Journal of Mineral Processing 157 (2016) 163–173

0.08
two-phase water-air system
0.06

Frequency
0.04

0.02

0
0 0.5 1 1.5 2 2.5 3
Bubble size (mm)

0.15

slurry system
0.1
Frequency

0.05

0
0 0.5 1 1.5 2 2.5 3
Bubble size (mm)

(a)

(b)
Fig. 6. (a) Modification of the bubble size distribution following the introduction of solids (4 wt%): creation of very small and large bubbles. (b) Images showing very fine and large bubbles
generated at 4% solids due to coalescence-induced break-up (5 ppm F150, Jg = 1 cm/s, and Jsw = 10.1 cm/s).

appear in the population, leading to a reduction of the gas hold-up. The coalescence during the generation stage since the water injected to
bubble coalescence induced by solid particles could occur during the the sparger does not contain solid particles. Therefore, the observed co-
generation process or progressively along the column. alescence would occur along the collection zone.
In the present study, a frit-sleeve sparger was used. The mechanism Coalescence-induced break-up occurs as one form of coalescence
of frit-sleeve spargers does not permit particles to induce bubble mechanism in which very fine bubbles are generated along with very
large ones (Tse et al., 2003; Quinn and Finch, 2012). This mechanism de-
creases the gas hold-up, as larger/faster rising bubbles are created.
On the other hand, it was suggested that in a three-phase fluidiza-
tion system, ‘sharp’ particle edges could penetrate and break some bub-
Table 2
Summary of the experimental plan.
ble surfaces. This would cause bubble to break-up into smaller ones,
thus leading to an increase in the gas hold-up. Yang et al. (2007) report-
Test # Frother concentration (ppm) Shearwater rate (cm/s) ed this phenomenon for large particles (larger than 1 cm). The exact ef-
1 5 10.1 fect of solids on bubble break-up is still not clear though. According to
2 5 20.2 Gandhi et al. (1999), the reduction in gas hold-up caused by solids addi-
3 5 40.4
tion would lead to a reduction in the bubble break-up rate. It must be
4 15 10.1
5 15 20.2 emphasized though that their experiments were conducted at high
6 15 40.4 slurry concentration (40 vol.%), and using very high gas rate (26 cm/s)
7 25 10.1 that are outside the typical range of operation.
8 25 20.2 The coexistence of both large and small bubbles was demonstrated
9 25 40.4
above in Fig. 6. Fig. 8 now illustrates the effect of solid particles on
A. Vazirizadeh et al. / International Journal of Mineral Processing 157 (2016) 163–173 169

2.2
53/75 m
75/106 m
106/150 m
53/150 m
2

1.8

d 32 [mm]

1.6

1.4

1.2

1
1 2 3 4 5 6 7 8 9
Test number

Fig. 7. The effect of the particle size on the bubble d32 at constant gas rate (1 cm/s).

bubble coalescence-induced break-up. Appendix A details the 3.2.2. Viscosity and bubble wake
corresponding hydrodynamic conditions. Results show that adding Based on the equation of Roscoe (1952), the slurry viscosity (μsl) can
solids (4% (w/w)) to the gas-water system and keeping the gas be calculated as
rate and frother concentration constant increased both the d10 and rel-
ative standard deviation of the distribution. As mentioned above, the μ sl ¼ μ l ð1−α Þ−2:5 ð4Þ
addition of hydrophobic particles thus promoted coalescence (larger
d10) and resulted in a more widespread distribution (larger relative where μl is the liquid viscosity, and α is the volume fraction of solids in
standard deviation), thus supporting the coalescence-induced break- the slurry.
up hypothesis. Increased slurry viscosity through the addition of solid particles
can stabilize the bubble wake. This, in turn leads to an increase of the

0.5
slurry system
two-phase water-air system
0.45

0.4
standard deviation [%]

0.35

0.3

0.25

0.2

0.15

0.1
0.55 0.6 0.65 0.7 0.75 0.8 0.85 0.9 0.95 1 1.05
d 10 [mm]

Fig. 8. The effect of solids on the parameters of the bubble size distribution (Jg = 1 cm/s, 25 ppm F150).
170 A. Vazirizadeh et al. / International Journal of Mineral Processing 157 (2016) 163–173

bubble velocity, and thus a reduction of the gas hold-up (Banisi et al., Table 3
1995). Measured gas hold-up and d32 for gas-water and gas-slurry systems (Jg = 1 cm/s,
Jsw = 40.4 cm/s, 25 ppm F150).
In the present work though (4% w/w solids or α = 1.5% v/v), μsl /
μl = 1.038, and the effect of viscosity on the gas hold-up is unlikely to Particle class (μm) εg (%) d32 (mm)
be dominant. However, it should be considered for higher solids frac- Two-phase – 12.95 0.63
tions and with this regard, a significant effect was reported by Banisi Three-phase 53/75 12.48 1.17
et al. (1995) for α N 5%.
3.2.3. Radial gas hold-up and flow profiles
Using a laboratory flotation column (4.47 m height and 10.18 cm in di-
ameter), Banisi et al. (1995) observed that the flat radial gas hold-up
and gas-slurry systems. The slurry system reports roughly the same
in the water-gas system changes to non-uniform profiles in the slurry-
gas hold-up in spite of a d32 value almost twice larger than that of
gas system (10% v/v calcite). They reported a significant difference be-
the gas-water, thus supporting the role of bubble loading on the
tween the distribution parameters of the gas hold-up in the gas-water
gas hold-up.
and gas-slurry systems due to a non-uniform profile using drift flux anal-
In addition to loading, it is possible that the bubble shape also plays a
ysis. This mechanism could explain the lower gas hold-up in the slurry
role in the difference of rise velocity, images taken revealing that larger
system.
bubbles tended to be less spherical than smaller ones (see Fig. 6). It is
The radial gas hold-up was not evaluated in the current study. How-
well documented that spherical bubbles exhibit lower rise velocities
ever, more recently Rabha et al. (2013) obtained contradictory experi-
than flattened ones (Maldonado et al., 2013) at a given frother concen-
mental results using a similar set-up (7 cm diameter and 150 cm
tration. In other words, the shape factor should further decrease the gas
height), and showing that the radial gas hold-up remained flat follow-
hold-up for large d32 values, which is not corroborated by the experi-
ing the addition of 5% solids, but decreased to a lower value. The mate-
mental results.
rial used was 100 μm spherical glass particle exhibiting a specific gravity
of 2500 kg/m3.
It must be emphasized that Banisi et al. (1995) applied the general
4. Discussion
drift-flux analysis method to study the radial hold-up, which is not
very accurate compared to the ultrafast electron beam X-ray tomogra-
Among the presented mechanisms, coalescence (and coales-
phy used by Rabha et al. (2013).
cence/induced break-up) and bubble loading can be used to explain
the variations of gas hold-up with the bubble size for two and three-
3.2.4. Bubble loading phase systems as depicted in Fig. 10. The first mechanism accounts
The added weight following hydrophobic particle attachment is for decreasing gas hold-up and the second one, for the opposite
known to reduce the bubble velocity and consequently increase the effect.
gas hold-up (Uribe-Salas et al., 2003). In the present case, particles By manipulating the shearwater rate, different bubble size distribu-
of talc attached on the bubble surface and were collected at the top tions can be generated while the gas rate and frother concentration re-
of the column. Fig. 9 demonstrates how talc particles readily at- main constant (25 ppm F150, Jg = 1 cm/s). As can be observed in Fig. 10,
tached to the bubble surfaces, and even accumulated over multiple different combinations of gas hold-up and bubble sizes were thus pro-
layers. The bubble loading effect can explain why the gas hold-up duced. Without great surprise, increasing the bubble size reduced the
was not reduced while increasing the bubble size. This is clearly gas hold-up in the two-phase operation.
demonstrated in Table 3, which compares test results in gas-water Adding solids produced coarser and more spread out bubble size dis-
tributions (see Fig. 6) due to coalescence. In most of the cases, this led to
a reduction of the gas hold-up that can be explained by the dominant ef-
fect of the bubble size on the rise velocity. It is worth mentioning that
shifting the d32 and increasing the standard deviation of the distribution
do not change the general shape of bubble size histogram, which can
still be represented by lognormal probability density function in this
case.
On the other hand, comparing the obtained results of the gas
hold-up in three-phase and two-phase systems also revealed that
in a few cases, notwithstanding coarser bubble size distributions,
the addition of solids led to higher gas hold-up values. The in-
creased weight of bubbles explains this behavior – in spite of the ef-
fect bubble shape favoring faster rise velocity for less spherical
bubbles –, i.e. the bubble rise velocity decreases as bubble loading
increases.
Results from the test conducted at different frother concentra-
tions in the presence of solids lead to the same conclusion:
coalescence and bubble–particle attachment are the main mech-
anisms explaining how the presence of solids affect the ties be-
tween the bubble size and gas hold-up. Reducing the frother
concentration below the CCC value revealed the effect of having
larger bubbles in the swarm. At the same time, decreasing the
number of bubbles due to coalescence increases bubble loading,
and thus decreases the rise velocity. Fig. 11 depicts how hydro-
phobic particles affect the correlation between the gas hold-up
and bubble size for three different frother concentrations. It is
Fig. 9. Loaded bubbles. worth noting that
A. Vazirizadeh et al. / International Journal of Mineral Processing 157 (2016) 163–173 171

0.25

0.2 Minimum bubble Minimum bubble


coalescence in two-phase coalescence &
0.15
maximum bubble
Frequency

0.1 loading in three-phase


0.25
0.05

0.2
0
0 10 20 30 40 50 60
Bubble size [mm]

0.15

Frequency
16 0.1

0.05

0
0 10 20 30 40 50 60 70
Bubble size [mm]

14

12
[%]

10
g

two-phase water-air system


slurry system
4
0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4 1.5
d32 [mm]

0.25

Medium bubble coalescence 0.2

& medium bubble loading in


three-phase 0.15
Frequency

0.1

0.05

0
0 10 20 30 40 50 60 70
Bubble size [mm]

Fig. 10. Bubble size and gas hold-up correlation in two and three-phase system.

1) εg ranging between 6 and 11% can be obtained for d32 ~1.3 mm, and 5. Conclusion
2) d32 ranging between 1.4 and 1.9 mm can produce εg ~5.5%.
Experimental results presented in this paper allowed studying
Very similar bubble size distributions – and gas hold-up – were mea- the effect of solid particles on bubble size distribution and gas
sured at 15 and 25 ppm as in both cases the concentration is above the hold-up, but also on the relationship between these two hydrody-
CCC. namic characteristics jointly. Different mechanisms introduced
172 A. Vazirizadeh et al. / International Journal of Mineral Processing 157 (2016) 163–173

0.25

0.2
Minimum bubble
coalescence &
0.15

maximum bubble
Frequency

0.1 loading in three-phase


0.05

0
0 10 20 30 40 50 60 70
Bubble size [mm]

14
25 ppm
15 ppm
5 ppm
12

0.25

Maximum bubble 0.2

10
coalescence &
0.15

medium bubble

Frequency
loading in three-phase 0.1
[%]

8 0.05
g

0
0 10 20 30 40 50 60 70
Bubble size [mm]

0.25

0.2

0.15
Frequency

0.1

1.2 1.4 1.6 1.8 2 2.2


0.05
d32 [mm]
0
0 10 20 30 40 50 60 70 0.25
Bubble size [mm]

Maximum bubble coalescence &


Medium bubble minimum bubble loading in
0.2

coalescence & medium three-phase 0.15


Frequency

bubble loading in three-


phase
0.1

0.05

0
0 5 10 15 20 25 30 35 40 45 50 55 60
Bubble size [mm]

Fig. 11. Bubble size and gas hold-up correlation in three-phase system for three frother concentrations.

from the literature were analyzed leading to the following 3. resulting effect: increasing the bubble size will not lead to the reduc-
conclusions: tion of the gas hold-up if the effect of the bubble loading outweighs
that of the bubble size on the rise velocity.
1. surface interactions: the effect of solid particles on the bubble size
distribution can be explained by coalescence through (1) the bridg- Characterizing the role of each one of these two mechanisms to de-
ing effect of solid particles and/or (2) the adsorption of frother by hy- termine the gas hold-up remains to be investigated.
drophobic surfaces;
2. bubble weight and volume: the effect of hydrophobic parti- Acknowledgements
cles on the gas hold-up is determined by a combination
of two antagonistic mechanisms: (1) bubble loading, which The authors acknowledge the financial support of NSERC grant # CRDPJ
reduces the rise velocity and (2) bubble coalescence, which 376316 – 2008 sponsored by COREM, Agnico-Eagle Mines – LaRonde
increases it; Division, and Glencore – Strathcona Mill (formerly Xstrata Nickel).
A. Vazirizadeh et al. / International Journal of Mineral Processing 157 (2016) 163–173 173

Appendix A. Experimental hydrodynamic conditions

Table A.1 provides the hydrodynamic conditions of the tests conducted with the water-air two-phase system. Table A.1 Hydrodynamic conditions for two-phase tests.

εg (%) d32 (mm) d10 (mm) σ2 σ Sb (1/s)

9.02 0.96 0.71 0.07 0.27 62.79


9.97 0.77 0.62 0.04 0.19 78.41
10.44 0.73 0.60 0.04 0.19 82.72
10.72 0.71 0.60 0.03 0.16 84.50
11.36 0.70 0.58 0.03 0.17 85.30
11.70 0.65 0.58 0.02 0.14 91.83
12.73 0.67 0.57 0.02 0.14 89.83
12.65 0.64 0.56 0.02 0.14 93.66
12.95 0.64 0.56 0.02 0.14 94.40
12.65 0.64 0.55 0.02 0.13 93.75
13.39 0.62 0.56 0.01 0.12 97.38
14.52 0.62 0.56 0.01 0.12 97.09
15.39 0.62 0.56 0.01 0.11 96.56

Table A.2 provides the hydrodynamic conditions of the tests conducted with slurry. Table A.2 Hydrodynamic conditions for three-phase tests.

εg (%) d32 (mm) d10 (mm) σ2 σ Sb (1/s)

7.10 1.36 0.97 0.17 0.41 44.23


9.85 1.21 0.86 0.12 0.35 49.71
12.48 1.17 0.85 0.11 0.34 51.27
6.85 1.35 0.96 0.17 0.41 44.32
9.68 1.26 0.92 0.13 0.37 47.66
10.86 1.30 0.92 0.15 0.39 46.06
4.58 1.35 0.95 0.17 0.42 44.57
7.93 1.24 0.95 0.13 0.36 48.25
10.46 1.18 0.91 0.11 0.33 50.93
5.11 1.46 1.01 0.21 0.46 41.22
7.85 1.33 0.99 0.16 0.40 45.19
9.34 1.31 0.96 0.15 0.39 45.73

References Maldonado, M., Desbiens, A., del Villar, Poulin R., Riquelme, A., 2015. Automatic control of
bubble size in a laboratory flotation column. Int. J. Miner. Process. 141, 27–33.
Banisi, S., Finch, J., Laplante, A., Weber, M., 1995. Effect of solid particles on gas holdup in O'Connor, C.T., Randall, E.W., Goodall, C.M., 1990. Measurement of the effects of physical
flotation columns—II. Investigation of mechanisms of gas holdup reduction in pres- and chemical variables on bubble size. Int. J. Miner. Process. 28 (1–2), 139–149.
ence of solids. Chem. Eng. Sci. 50 (14), 2335–2342. Quinn, J.J., Finch, J.A., 2012. On the origin of bi-modal bubble size distributions in the ab-
Dippenaar, A., 1982. Destabilization of froth by solids - 1. The mechanism of film rupture. sence of frother. Miner. Eng. 36–38 (0), 237–241.
Int. J. Miner. Process. 9 (1), 1–14. Rabha, S., Schubert, M., Wagner, M., Lucas, D., Hampel, U., 2013. Bubble size and radial gas
Finch, J., Dobby, G., 1990. Column Flotation. Pergamon Press . hold-up distributions in a slurry bubble column using ultrafast electron beam X-ray
Finch, J.A., Wills, B.A., 2015. Wills' Mineral Processing Technology: An Introduction to the tomography. AICHE J. 59, 1709–1722.
Practical Aspects of Ore Treatment and Mineral Recovery. Butterworth-Heinemann . Reese, J., Jiang, P., Fan, L.-S., 1996. Bubble characteristics in three-phase systems used for
Finch, J.A., Nesset, J.E., Acuna, C., 2008. Role of frother on bubble production and behaviour pulp and paper processing. Chem. Eng. Sci. 51 (10), 2501–2510.
in flotation. Miner. Eng. 21 (12–14), 949–957. Riquelme, A., Desbiens, A., Bouchard, J., del Villar, R., 2013a. Parameterization of bubble
Gandhi, B., Prakash, A., Bergougnou, M.A., 1999. Hydrodynamic behavior of slurry bubble size distribution in flotation columns. 16th IFAC Symposium on Automation in Min-
column at high solids concentrations. Powder Technol. 103 (2), 80–94. ing, Minerals and Metal Processing, MMM 2013a, August 25–28, 2013 San Diego, CA,
Gomez, C.O., Cortés-López, F., Finch, J.A., 2003. Industrial testing of a gas holdup sensor for USA .
flotation systems. Miner. Eng. 16 (6), 493–501. Riquelme, A., Bouchard, J., Desbiens, A., del Villar, R., 2013b. Bubble detection in flotation
Grau, R.A., Heiskanen, K., 2005. Bubble size distribution in laboratory scale flotation cells. columns based on circular Hough transform. 23rd World Mining Congress, August
Miner. Eng. 18 (12), 1164–1172. 11–15, 2013, Montréal, Canada .
Hernandez-Aguilar, J.R., Coleman, R.G., Gomez, C.O., Finch, J.A., 2004. A comparison be- Roscoe, R., 1952. The viscosity of suspensions of rigid spheres. Br. J. Appl. Phys. 3 (8), 267.
tween capillary and imaging techniques for sizing bubbles in flotation systems. Tse, K., Martin, T., McFarlane, C., Nienow, A., 2003. Small bubble formation via a coales-
Miner. Eng. 17 (1), 53–61. cence dependent break-up mechanism. Chem. Eng. Sci. 58 (2), 275–286.
Kracht, W., Vallebuona, G., Casali, A., 2005. Rate constant modelling for batch flotation, as Tsutsumi, A., Nieh, J.-Y., Fan, L.-S., 1991. Particle wettability effects on bubble wake dy-
a function of gas dispersion properties. Miner. Eng. 18 (11), 1067–1076. namics in gas-liquid-solid fluidization. Chem. Eng. Sci. 46 (9), 2381–2384.
Kuan, S.H., 2009. The Effect of Solids on Gas Holdup, Bubble Size and Water Overflow Rate Uribe-Salas, A., de Lira-Gomez, P., Perez-Garibay, R., Nava-Alonso, F., Magallanes-
in Flotation. M.Sc. thesis, McGill University . Hernandez, L., Lara-Valenzuela, C., 2003. Overloading of gas bubbles in column flota-
Kuan, S.H., Finch, J.A., 2010. Impact of talc on pulp and froth properties in F150 and 1- tion of coarse particles and effect upon recovery. Int. J. Miner. Process. 71 (1–4),
pentanol frother systems. Miner. Eng. 23 (11), 1003–1009. 167–178.
Luo, X., Lee, D., Lau, R., Yang, G., Fan, L.S., 1999. Maximum stable bubble size and gas hold- Vazirizadeh, A., Bouchard, J., del Villar, R., 2015. On the relationship between hydrody-
up in high-pressure slurry bubble columns. AICHE J. 45 (4), 665–680. namic characteristics and the kinetics of column flotation. Part I: modeling the gas
Maldonado, M., Desbiens, A., del Villar, R., 2008a. An update on the estimation of the froth dispersion. Miner. Eng. 74, 207–215.
depth using conductivity measurements. Miner. Eng. 21 (12–14), 856–860. Vinnett, L., Contreras, F., Yianatos, J., 2012. Gas dispersion pattern in mechanical flotation
Maldonado, M., Desbiens, A., del Villar, R., Girgin, E., Gomez, C., 2008b. On-line estimation cells. Miner. Eng. 26, 80–85.
of bubble size distributions using Gaussian mixture models. V International Mineral Yang, G.Q., Du, B., Fan, L.S., 2007. Bubble formation and dynamics in gas–liquid–solid
Processing Seminar . fluidization—a review. Chem. Eng. Sci. 62 (1–2), 2–27.
Maldonado, M., Quinn, J.J., Gomez, C.O., Finch, J.A., 2013. An experimental study examin-
ing the relationship between bubble shape and rise velocity. Chem. Eng. Sci. 98, 7–11.

You might also like