Sunteți pe pagina 1din 10

Minerals Engineering 77 (2015) 121–130

Contents lists available at ScienceDirect

Minerals Engineering
journal homepage: www.elsevier.com/locate/mineng

The interaction of clay minerals with gypsum and its effects on


copper–gold flotation
Nestor Cruz a, Yongjun Peng a,b,⇑, Elaine Wightman a, Ning Xu c
a
Julius Kruttschnitt Mineral Research Centre, University of Queensland, Isles Road, Indooroopilly, Brisbane, QLD 4068, Australia
b
School of Chemical Engineering, University of Queensland, St. Lucia, Brisbane, QLD 4072, Australia
c
Minerals and Materials Science & Technology, University of South Australia, Mawson Lakes, SA 5095, Australia

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

Article history: The interaction of two clay minerals, kaolinite and bentonite with gypsum and its effects on the flotation
Received 28 January 2015 of a copper–gold ore was investigated in this study. It was found that bentonite increased the viscosity
Revised 10 March 2015 more than kaolinite when mixed with the copper–gold ore at low shear rates. The detrimental effect
Accepted 11 March 2015
of these clay minerals on flotation was attributed to the entrainment of clay particles when kaolinite
Available online 25 March 2015
was added to the ore and to a decrease in true flotation by bentonite. Bentonite formed a sponge-like
structure with predominant edge–edge (E–E) interactions which might affect hydrodynamics in the flota-
Keywords:
tion cell and have a detrimental effect on flotation recovery. Kaolinite did not form a particular network
Kaolinite
Sodium bentonite
structure and its aggregates mostly consisted of face–face (F–F) type associations which did not affect
Gypsum flotation hydrodynamics. The addition of gypsum to the ore–bentonite mixture inhibited the formation
Rheology of interconnected network structures. This led to lower viscosity values with flotation behaviour similar
Network structures to that of mixtures with kaolinite. In this case, there was an improvement in recovery, but the grade
decreased due to entrainment. The addition of gypsum to the ore–kaolinite mixture created aggregates
with long strings further enhancing particle entrainment with more mass transported to the froth.
Ó 2015 Elsevier Ltd. All rights reserved.

1. Introduction The factors influencing hydrodynamics are flotation cell charac-


teristics, impeller properties, and slurry properties such as density
Clay minerals can be detrimental to mineral flotation and high and rheology (Shabalala et al., 2011). In this paper the only factor
viscosity and high particle entrainment promoted by clay minerals that is changed is the slurry rheology by adding clay minerals.
are among the main contributing factors (Forbes et al., 2014; This increases pulp viscosity and may cause a high yield stress
Jorjani et al., 2011; Wang and Peng, 2014; Wei et al., 2013; and apparent viscosities in the flotation cell, which decreases the
Zhang and Peng, 2015). Incremental reagent consumption has also bubble-particle collision efficiency (Schubert, 2008; Xu et al.,
been suggested as a factor affecting flotation of ores with a high 2011) and changes gas dispersion parameters such as bubble size,
clay content (Connelly, 2011). Bentonite has a 2:1 layer structure gas hold-up, superficial gas velocity and bubble surface area flux
and is a swelling clay mineral while kaolinite does not swell and (Shabalala et al., 2011). It is reported that a high yield stress can
has a 1:1 structure. Both have a small particle size. Zhang and form a ‘‘cavern’’ around the impeller (Bakker et al., 2010) and this
Peng (2015) found that bentonite increased the viscosity of a cop- decreases bubble size and gas hold-up. When a ‘‘cavern’’ exists
per–gold ore slurry and decreased copper and gold flotation recov- small bubbles are formed in the impeller zone, but the dispersion
ery, while kaolinite affected the viscosity slightly with of these bubbles is poor in the cell (Shabalala et al., 2011). Patra
simultaneous high gangue entrainment. The decrease in mineral et al. (2012a, 2012b) found that network structures formed by
flotation recovery by bentonite may be via the modification of fibrous materials in flotation also interfered with the hydrody-
hydrodynamic conditions inside flotation cells, and this is related namic conditions and then decreased mineral flotation recovery.
to fluid flow which is mostly driven by the action of the impeller. Their studies suggest that network structures formed by bentonite
may have a similar effect on mineral flotation.
Wang and Peng (2014) studied the negative effect of clay
minerals in flotation through the transport of these particles to
⇑ Corresponding author at: School of Chemical Engineering, University of
the concentrate. They found that the degree of entrainment of clay
Queensland, St. Lucia, Brisbane, QLD 4072, Australia. Tel.: +61 7 3365 7156; fax:
+61 7 3365 3888. particles was high and also affected by the presence of electrolytes
E-mail address: yongjun.peng@uq.edu.au (Y. Peng). in the process water. The high particle entrainment in flotation in

http://dx.doi.org/10.1016/j.mineng.2015.03.010
0892-6875/Ó 2015 Elsevier Ltd. All rights reserved.
122 N. Cruz et al. / Minerals Engineering 77 (2015) 121–130

the presence of kaolinite was attributed to the formation of aggre- Particle size measurements were conducted with a Mastersizer
gates which entered flotation concentrates and then enhanced Microplus from Malvern Instruments. This equipment measures
froth stability (Wang and Peng, 2014). Electrolytes in saline water particle size distribution in very dilute solutions by using laser
contributed to the formation of aggregates that were entrained, but diffraction with a standard procedure to avoid particle agglomera-
when the clay content was too high, entrainment and true flotation tion. The P80 of both kaolinite and bentonite was about 14 lm, and
were affected due to high pulp viscosity which limited bubble and the P80 of gypsum was about 77 lm. Further size reduction for
particle mobility. In that case, edge–edge (E–E) network structures gypsum was difficult due to agglomeration that occurred during
were enhanced (Wang and Peng, 2014). pulverization. Quartz had an original P80 of about 52 lm, and it
Calcium bearing minerals can be present in ores and associate was pulverized to achieve a P80 of 14 lm. Its final particle size
with clay minerals. In the previous work, it was found that among distribution was very close to the size distribution of the clay
the calcium bearing minerals gypsum had the strongest rheological minerals.
interaction with bentonite and kaolinite (Cruz et al., 2013).
Gypsum acted as a potential source for Ca2+ and SO24 ions released
into the process water. It is known that multivalent cations are 2.2. Preparation of mineral slurries
more strongly attracted to clay particles than monovalent cations
(i.e. Na+ ions) (Lagaly and Dékány, 2013a). Ca2+ cations can increase The low-clay copper–gold ore was used to prepare five different
the viscosity of kaolinite slurries by attaching to the clay surfaces slurries by mixing it with known amounts of kaolinite (30 wt.%),
modifying the double layers and contributing to the formation of sodium bentonite (15 wt.%), kaolinite–gypsum (30–5 wt.%),
network structures (Abdi and Wild, 1993; Wild et al., 1993). In sodium bentonite–gypsum (15–5 wt.%), and quartz (30 wt.%). The
bentonite slurries a high concentration of Ca2+ cations can replace low-clay ore with no clay addition was used as a baseline. The
Na+ cations in the interlayer space deactivating sodium bentonite mixture with quartz served two purposes: (1) understanding the
such that it is no longer possible for hydration and swelling to effect of the dilution of copper and gold in the feed on the flotation,
occur (Alther, 1986; Bradshaw et al., 2013; Meer and Benson, and (2) comparing the effect from quartz and clay minerals. A
2007). It is also reported that the presence of gypsum and lime 30 wt.% quartz content was chosen since it was the highest clay
can increase water absorption and swelling pressure of bentonite content used. Although the particle size of the quartz is smaller
(Abdi and Wild, 1993; Wild et al., 1993). The Ca2+ from lime and than the particle size of the low-clay ore, it can still provide infor-
gypsum and SO24 anions play a role on this effect. mation about the effect on recovery when feed grade is changed
In this study, the interaction of clay minerals with gypsum in during the mixture preparation. The solid concentration for all
the formation of network structures and its effects on copper–gold the slurries was maintained at 30 wt.% (about 14.6 vol.%) and the
flotation were investigated. total weight of the mixtures was 1 kg for all cases. The slurries
were prepared using Brisbane tap water.
The low-clay ore was wet ground to a P80 of 106 lm at a
2. Experimental 60 wt.% solid concentration. Hydrated lime (800 g/t) and promoter
Aero 3894 (8 g/t) were added to the mill to mimic the plant scale
2.1. Materials processing. After grinding, the slurry was transferred to a 3 L
flotation cell where it was well mixed at 1500 RPM and a known
For the flotation experiments sodium isopropyl xanthate (SIPX), volume of slurry was removed. The same weight of solids that
and the promoter Aero 3894 were used as the collectors. Both was removed was replaced with quartz, clay mineral or a mixture
reagents were supplied by Cytec. The frother Polyfroth W22 was of clay mineral and gypsum. Water was added to make up the
supplied by Huntsman Performance Products, and it is low slurry to 30 wt.% for flotation. The removed slurry was filtered,
molecular weight polyoxyalkylene alkyl ether frother partially dried and weighted to confirm that a correct amount of solids
soluble in water. The pH in flotation was adjusted using AR grade was taken from the original slurry.
hydrated lime (Ca(OH)2). These reagents are used in the industrial After adding the quartz, clay mineral or the mixture of clay min-
flotation of the copper–gold ore examined in this study. eral and gypsum to the flotation cell, stirring continued at 900 RPM
The gypsum (calcium sulphate dihydrate) that was added to the for 30 min when adding quartz or kaolinite and for 45 min when
ore–clay mineral mixtures had a high purity (P99%), and was pur- adding bentonite. When adding bentonite to the ore, agglomerates
chased from Sigma–Aldrich. Kaolinite Q38, sodium bentonite, and were formed and more mixing time was required. It was also
quartz were supplied by Sibelco Group, Australia. Q38 is a dry observed that after 30 or 45 min the rheology measurements
milled kaolinite with a surface area of 26 m2/g. Quantitative XRD became consistent for the mixtures with kaolinite or bentonite.
analysis showed that the composition of the kaolinite sample Once the slurry was mixed, the pH was adjusted to 10 by the
was 85 wt.% kaolinite, 4 wt.% quartz and 11 wt.% muscovite, while addition of hydrated lime.
the composition of the sodium bentonite sample was 63 wt.%
montmorillonite, 25 wt.% albite and 12 wt.% quartz. The relatively
low-crystallinity of Q38 kaolinite with Hinckley crystallinity index 2.3. Rheology measurements
(IH) of 0.5, derived from the XRD pattern (Plancon et al., 1988), was
chosen in our experiments because it is close to the IH of the A 20 mL syringe was used to draw a sample from the flotation
kaolinite in the copper–gold ore (0.6). cell after the addition of flotation reagents. This sample was imme-
The copper–gold ore was obtained from one of the sponsors. diately transferred to the cup of an Anton Paar DSR 301 rheometer.
This was considered a low-clay ore. Quantitative XRD analysis indi- This cup was part of a Couette geometry (bob and cup) that was
cated that the mineral composition of the ore was 3 wt.% kaolinite, installed for this equipment. A rheogram was produced in 35 s to
1 wt.% pyrite, 28 wt.% quartz, 10 wt.% muscovite, 50 wt.% albite reduce the effect of particle settling. The tests were done at ambi-
and 8 wt.% clinochlore. The head assays of gold and copper were ent temperature (23 °C), and the shear rate for the rheograms was
0.46 ppm and 0.55 wt.%, respectively. 1 kg crushed ore was ground between 0.1 and 350 s 1, but the shear rate of the rheograms
to a P80 of 106 lm before preparing the different mixtures shown in this paper was from 0.1 to 222 s 1. For some slurries with
with clay minerals and gypsum. By following this procedure the low viscosities, turbulence was created inside the Couette geome-
valuable mineral liberation was similar for all the experiments. try at shear rate values more than 222 s 1.
N. Cruz et al. / Minerals Engineering 77 (2015) 121–130 123

2.4. Flotation experiments shows the mass recovery as a function of water recovery from
these tests. As can be seen, flotation of the ore or baseline resulted
Four concentrates were collected after 1, 3, 7, and 17 min of in a 5% mass recovery and 31% water recovery at the completion of
flotation. During flotation the impeller speed was set at 800 RPM, flotation. The flotation of the ore and its mixture with 30 wt.%
and the air flow rate was 8 L/min. The pH was adjusted to 10 with quartz produced very similar mass and water recoveries. The addi-
hydrated lime before collecting the concentrates and the collector tion of 15% bentonite reduced the water recovery significantly to
SIPX and frother Polyfroth W22 were added prior to collecting the about 9%. This was because a froth phase was almost absent during
first and third concentrates. The collector dosage was 6 and 4 g/t the flotation of this mixture. The addition of 30% kaolinite
before the first and third concentrate respectively. 15 g/t frother increased water recovery and mass recovery significantly to 38%
was added before the first and third concentrates. By using this and 24%, respectively. The addition of 5 wt.% gypsum to the
procedure, a similar flotation performance from the flotation of ore–bentonite mixture produced much higher water and mass
the low-clay ore was achieved in the laboratory and plant. recoveries due to the formation of a more stable froth phase. The
addition of 5 wt.% gypsum to the ore–kaolinite mixture also
2.5. Cryo-SEM analyses increased water and mass recoveries, but to a lower degree than
the gypsum addition to the ore–bentonite mixture.
The Cryo-SEM technique has been used in many applications to Figs. 2 and 3 show copper grade as a function of copper recov-
avoid changes in structure induced by vacuum drying or freeze- ery, and gold grade as a function of gold recovery, respectively,
drying since the water is vitrified without crystallisation as ice from the flotation of the ore and its mixtures. The main observation
(Battersby et al., 1994). Two 6 mm long copper tubes with outer- was the poor flotation outcome that resulted from the mixture
diameter 4 mm and inner-diameter 3 mm were joined together with bentonite and its change after gypsum addition. The baseline
by superglue to hold liquid samples. The samples collected from flotation of the ore resulted in a copper recovery of 92% at a grade
pulp and the froth of first concentrate were transferred into the of 10% copper (Fig. 2), and 81% gold recovery at a grade of 7 ppm
copper tube using a wide-bore pipette. The copper tube with the gold (Fig. 3). These copper and gold recoveries were almost the
sample inside was immediately sealed with dental wax, placed in same as those obtained from the flotation of the ore–quartz
a container and plunged into liquid nitrogen. Each sample was mixture, but the addition of quartz caused a decrease in grade to
preserved in liquid nitrogen until transferred to the sample 6% for copper and 4 ppm for gold. It seems that in this work the
preparation chamber of the PHILIPS XL30 field emission gun decrease in feed grade did not affect copper or gold recovery. The
scanning electron microscope (FESEM) equipped with Oxford effect on the concentrate grade after the addition of quartz may
Cryo-transfer and fracture stage. The top tube of the two glue- be due to the higher entrainment of smaller quartz particle, which
jointed tubes was then knocked off by a metal knife on the fracture was examined below.
stage to expose a fresh sample surface inside the tube. The sample Although copper and gold grades from the flotation of the ore–
temperature was raised to 175 K to sublime vitrified water at the bentonite mixture were close to the baseline flotation, copper and
rate of 6 nm/s for 2 min with an estimated 720 nm depth of gold recoveries were much lower (Figs. 2 and 3). The addition of
vitrified water sublimated. The sample was eventually coated by gypsum to the bentonite–ore system changed the flotation
platinum plasma for 3 min to form a 3 nm thick platinum coating response significantly by improving the copper recovery from
to avoid charging during the imaging process. FESEM was operated 83% to 93% and gold recovery from 64% to 85%. Meanwhile, the
at 10 kV voltages coupled with Energy-dispersive X-ray spec- addition of gypsum to the ore–bentonite mixture decreased copper
troscopy (EDS) to identify the elemental composition of particles. and gold grades throughout the flotation (Figs. 2 and 3). For exam-
A preliminary report on the application of this technique to studies ple, copper grade went down from 8% to 2%, and the gold grade
of kaolinite aggregation in pulp, froth and concentrate phases was from about 5 ppm to 1 ppm in the end of flotation. When gypsum
published by Xu et al. (2014). was added to the ore–bentonite mixture, it was observed that the
froth became more abundant and stable as supported by the higher
mass–water recovery curve in Fig. 1.
3. Results
The addition of kaolinite to the ore did not decrease copper and
gold recoveries from the flotation as the addition of quartz, but did
3.1. Flotation
decrease copper and gold grades to about 2%, and 1 ppm as shown
in Figs. 2 and 3, respectively more than the addition of quartz. The
Six flotation tests were conducted using the copper–gold ore
addition of gypsum to the ore–kaolinite mixture did not further
and its mixtures with quartz, clay minerals and gypsum. Fig. 1
affect copper and gold flotation significantly.
100 wt.% ore
30 wt.% kaolinite - 70 wt.% ore 100 wt.% ore
30 wt.% kaolinite - 5 wt.% gypsum - 65 wt.% ore 30 wt.% kaolinite - 70 wt.% ore
15 wt.% bentonite - 85 wt.% ore 30 wt.% kaolinite - 5 wt.% gypsum - 65 wt.% ore
35 15 wt.% bentonite - 5 wt.% gypsum - 80 wt.% ore 15 wt.% bentonite - 85 wt.% ore
30.0
30 wt.% quartz - 70 wt.% ore 15 wt.% bentonite - 5 wt.% gypsum - 80 wt.% ore
30 30 wt.% quartz - 70 wt.% ore
25.0
Mass recovery (%)

25
Copper grade (%)

20.0
20
15.0
15
10.0
10

5 5.00

0 0.00
0 5 10 15 20 25 30 35 40 45 50 0.00 10.0 20.0 30.0 40.0 50.0 60.0 70.0 80.0 90.0 100.0
Water recovery (%) Copper recovery (%)

Fig. 1. Mass recovery as a function of water recovery from the flotation of the low- Fig. 2. Copper grade as a function of copper recovery from the flotation of the ore
clay ore and its mixtures with quartz, clay minerals and gypsum. and its mixtures with quartz, clay minerals and gypsum.
124 N. Cruz et al. / Minerals Engineering 77 (2015) 121–130

100 wt.% ore 100 wt.% ore


30 wt.% kaolinite - 70 wt.% ore 30 wt.% kaolinite - 70 wt.% ore
30 wt.% kaolinite - 5 wt.% gypsum - 65 wt.% ore 30 wt% kaolinite - 5 wt.% gypsum - 65 wt.% ore

Copper recovery by entrainment (%)


20.0 15 wt.% bentonite - 85 wt.% ore 5.0 15 wt.% bentonite - 85 wt.% ore
15 wt.% bentonite - 5 wt.% gypsum - 80 wt.% ore 15 wt.% bentonite - 5 wt.% gypsum - 80 wt.% ore
30 wt.% quartz - 70 wt.% ore 30 wt.% quartz - 70 wt.% ore
4.0
15.0
Gold grade (ppm)

3.0
10.0
2.0

5.00 1.0

0.0
0.00 0 2 4 6 8 10 12 14 16 18
0.00 10.0 20.0 30.0 40.0 50.0 60.0 70.0 80.0 90.0 10.00
Gold recovery (%) Flotaon me (min)

Fig. 3. Gold grade as a function of gold recovery from the flotation of the ore and its Fig. 5. Copper recovery by entrainment from the flotation of the ore and its
mixtures with quartz, clay minerals and gypsum. mixtures with quartz, clay minerals and gypsum.

The increase in recovery can be attributed to recovery by both From these flotation results it can be inferred that the presence
entrainment and true flotation. The Savassi equation (Bıçak et al., of bentonite reduced true flotation, but the presence of gypsum
2012; Savassi et al., 1998) was used to calculate copper recovery caused bentonite to behave more like kaolinite. Kaolinite increased
by true flotation and results are shown in Fig. 4. The overall copper the recovery by entrainment as a result of the creation of a more
recovery by true flotation from the flotation of the ore and its stable froth which is supported by the high mass–water recovery
mixture with quartz was similar (about 92%). The true flotation curve (Fig. 1). This entrainment increased the overall recovery,
recoveries for the mixtures with clay minerals and gypsum were but diluted the concentrate grade (Figs. 2 and 3). The addition of
close to the baseline (about 91%) except for the flotation of gypsum to the kaolinite–ore system promoted additional entrain-
ore–bentonite mixture that produced much lower copper recovery ment, further increasing the recovery but decreasing the grade
by true flotation (82%). However, by examining copper recovery (Figs. 2–5).
from the first concentrate in Fig. 4, all the mixtures apart from
the bentonite mixture, produced higher copper recovery than the
3.2. Rheology measurements
baseline case. It seems that the addition of kaolinite to the ore
increased copper recovery by true flotation marginally, but the
The rheology measurements of the ore and its mixtures with
addition of bentonite decreased copper recovery by true flotation
quartz, clay minerals and gypsum at pH 10 are shown in Fig. 6.
significantly. The addition of gypsum with bentonite altered the
The ore and its mixture with quartz had the lowest shear stress
deleterious effect of bentonite and improved copper flotation by
values and lowest viscosities. The addition of 30 wt.% kaolinite or
true flotation.
15 wt.% sodium bentonite to the ore increased the viscosity as
Fig. 5 shows the copper recovery by entrainment. The flotation
expected (Cruz et al., 2013). The addition of gypsum reduced the
of the ore produced very low copper recovery by entrainment (only
rheological properties of bentonite slightly, but had little effect
0.5%). The addition of quartz to the ore only increased the copper
on the rheological properties of kaolinite as shown in Fig. 6. Due
recovery by entrainment slightly. The flotation of the ore–kaolinite
to the swelling property of bentonite, 15 wt.% bentonite increased
mixture produced a much higher overall copper recovery by
the viscosity to the same degree as 30 wt.% kaolinite when mixed
entrainment (3.2%) than the flotation of the ore–bentonite mixture
with the ore.
(1.2%). The addition of gypsum significantly changed this recovery
Adjusting the pH with hydrated lime added Ca2+ cations to the
especially in the case of the ore–bentonite mixture where the
slurry, and gypsum contributed to additional Ca2+ cations that
recovery by entrainment was increased to 3.0%. In the mixture
should cause kaolinite particles to form agglomerates increasing
with kaolinite, gypsum increased the recovery by entrainment to
the viscosity. However, this increase was not observed when gyp-
3.8%.
sum was added to the ore–kaolinite mixture. The unexpected

100 wt.% ore


100.0 30 wt % kaolinite - 70 wt.% ore
30 wt. % kaolinite - 5 wt.% gypsum - 65 wt.% ore
Copper recovery by true flotaon (%)

2.0 15 wt.% bentonite - 85 wt.% ore


15 wt.% bentonite - 5 wt.% gypsum - 80 wt.% ore
80.0 30 wt.% quartz - 70 wt.% ore

1.5
Shear stress (Pa)

60.0

1.0
100 wt.% ore
40.0 30 wt.% kaolinite - 70 wt.% ore
30 wt% kaolinite - 5 wt.% gypsum - 65 wt.% ore
15 wt.% bentonite - 85 wt.% ore 0.5
20.0 15 wt.% bentonite - 5 wt.% gypsum - 80 wt.% ore
30 wt.% quartz - 70 wt.% ore

0.0
0.00 0 50 100 150 200 250
0 2 4 6 8 10 12 14 16 18 Shear rate (s-1)
Flotaon me (min)
Fig. 6. Rheograms of the ore and its mixtures with quartz, clay minerals and
Fig. 4. Copper recovery by true flotation from the flotation of the ore and its gypsum: total solids concentration 30 wt.% and pH adjusted to 10 with hydrated
mixtures with quartz, clay minerals and gypsum. lime (Ca(OH)2).
N. Cruz et al. / Minerals Engineering 77 (2015) 121–130 125

rheology response was further examined by Cryo-SEM analyses in 1.0


the next section. In contrast gypsum decreased the viscosity of the 0.9

Apparent yield stress (Pa)


ore–bentonite mixture by either inhibiting the formation of the 0.8
bentonite network structures or reducing its swelling properties. 0.7

At a first glance the rheograms in Fig. 6 did not show any major 0.6

difference between the ore and its mixtures with clay minerals and 0.5

gypsum. The rheograms for ore–kaolinite and ore–bentonite 0.4

mixtures were similar, but the flotation response for these two 0.3

mixtures was very different. One difference between the 0.2

ore–bentonite mixture and other mixtures was the higher shear 0.1

stress values in the shear rate range between 0.1 and 50 s 1 0.0
Low-clay ore and Kaolinite 30 wt. % - Kaolinite 30 wt % - Bentonite 15 wt.% Bentonite 15 wt.%
(Fig. 6). This suggests that for rheology it may be better to compare Quartz 30 wt.% - Gypsum 5 wt.% - ore 70 wt.% - gypsum 5 wt.% - - ore 85 wt.%
Ore 70 wt.% ore 65 wt.% oren 80 wt.%
the yield stress regions of the rheograms or the shear stress values mixture
at low shear rates. Shear rates in a flotation cell vary from high to
low values depending on the proximity of the slurry to the Fig. 8. Calculated apparent yield stresses for the ore and its mixtures with quartz,
clay minerals and gypsum using Herschel–Buckley model.
flotation cell impeller and the rheology of the slurry (Bakker
et al., 2009). This causes the ore–clay mixtures which are
non-Newtonian, to have different shear stresses and viscosities at rheology tests suggest that the high viscosity could be one con-
different locations in the flotation cell. It has been demonstrated tributing factor. In this section Cryo-SEM was used to understand
that there is a considerable variation of the shear rate distribution the types of aggregates and network structures formed by clay
inside the flotation cell with the highest values close to the minerals and their interactions with gypsum.
impeller (Bakker et al., 2009). The other areas in the flotation cell Cryo-SEM images in Figs. 9 and 10 compare the type of aggre-
may have relatively small shear rate values that can be even close gates and network structures present in the pulp of ore–bentonite
to zero (Anderson, 2008). and ore–kaolinite mixtures at different magnifications. It is clear
Fig. 7 shows the relationship between apparent viscosity and that there are differences in the kaolinite and bentonite particle
shear stress. The ore–bentonite mixture had much higher apparent aggregates at the same pH. Figs. 9a and 10a show that the ben-
viscosity values than the other mixtures in the shear stress range tonite created a honeycomb-like network structure. This structure
between 0.5 and 0.9 Pa. In this shear stress range the shear rate appears to be made of many interconnected cells with pore size at
for this mixture is between 0.25 and 6 s 1 (Fig. 6). This means that about 5–10 lm in diameter where E–E and E–F aggregates were
for the zones in the flotation cell with shear rates in that range, the predominant. Figs. 9b and 10b show that the kaolinite platelets
apparent viscosity values are greater when a 15 wt.% bentonite is associated mainly in the mode of F–F and E–E as dense elongated
present in the mixture with the ore. strings and the resulting aggregates were joined to each other
Figs. 6 and 7 suggest that the Herschel–Bulkley model may be commonly in an E–F manner to form loose clumps. No honey-
more appropriate to calculate the yield stress for the slurries used comb-like network structure was observed in the pulp of ore–
in this work since this model fits the yield stress region in the rheo- kaolinite mixture.
grams well. Fig. 8 shows the calculated apparent yield stresses for Since a greater proportion of kaolinite was added compared to
the ore and its mixtures with clay minerals and gypsum using the sodium bentonite, 30 and 15 wt.% respectively, the kaolinite aggre-
Herschel–Buckley model. Results show that there is a difference in gates look more crowded with larger aggregates. Despite this, the
yield stress between the ore–bentonite mixture (0.9 Pa) and the slurry containing sodium bentonite had higher shear stress and
ore–kaolinite mixture (0.5 Pa). The trend in yield stress in Fig. 8 viscosity values at low shear rates as it was explained in the rheol-
is well correlated to flotation outcomes and the higher the yield ogy section. This suggests that the network structure made by ben-
stress, the lower the flotation mass and metal recovery. tonite particles provided much more resistance to flow than the
loose kaolinite aggregates.
Wang and Peng (2014) explained that the entrainment of clay
3.3. Cryo-SEM analysis particles in flotation depended on the type of clay particle associa-
tion, and the recovery of clay minerals increased if they entered the
The flotation results confirm that both kaolinite and bentonite froth as aggregates increasing froth stability. They stated that
have a detrimental effect on copper and gold flotation, and edge–edge (E–E) associations of clay platelets gave higher viscosi-
ties resulting in poor flotation. Face–face (F–F) aggregates are more
compact with denser structures giving lower viscosities. In this
2500 100 wt.% ore current study, this difference in viscosity caused by the E–E and
30 wt % kaolinite - 70 wt.% ore
30 wt. % kaolinite - 5 wt.% gypsum - 65 wt.% ore F–F aggregates was observed in the mixtures with bentonite (E–E
15 wt.% bentonite - 85 wt.% ore
predominant) and kaolinite (mostly F–F associations).
Apparent viscosity (mPa.s)

2000 15 wt.% bentonite - 5 wt.% gypsum - 80 wt.% ore


30 wt.% quartz - 70 wt.% ore Figs. 11 and 12 show the Cryo-SEM images of the network
1500 structure formed by ore–bentonite and ore–bentonite–gypsum in
flotation pulp at different magnifications. As can be seen, when
1000 gypsum was added to the ore–bentonite mixture, the extensive
E–E connected long strings were formed without any E–F associa-
500 tions and the bentonite aggregates were thicker due to more F–F
associations as compared to the ore–bentonite mixture
0 (Figs. 11b and 12b). These observed changes in network structure
0 0.5 1 1.5 2 2.5
agree with the findings from Morris and Zbik _ (2009) who added
Shear stress (Pa) 0.05 M CaCl2 to a 3 wt.% bentonite suspension. Although the
5 wt.% gypsum in the slurries used in this work gave approxi-
Fig. 7. Apparent viscosity as a function of shear stress (Pa) for the ore and its
mixtures with quartz, clay minerals and gypsum: pH adjusted to 10 with hydrated mately a 0.016 M concentration of Ca2+ cations, the change in the
lime (Ca(OH)2). network structures were still evident. The presence or absence of
126 N. Cruz et al. / Minerals Engineering 77 (2015) 121–130

Fig. 9. Comparison of the sodium bentonite network structures (a) and kaolinite aggregates (b) in the ore–clay mixtures in flotation pulp: sodium bentonite and kaolinite
concentrations are 15 wt.% and 30 wt.% respectively; 1000 magnification.

Fig. 10. Comparison of the sodium bentonite network structures (a) and kaolinite aggregates (b) in the ore–clay mixtures in flotation pulp: sodium bentonite and kaolinite
concentrations are 15 wt.% and 30 wt.% respectively; 4000 magnification.

Fig. 11. Cryo-SEM images showing the network structure formed by ore–bentonite (a) and ore–bentonite–gypsum (b) mixtures in flotation pulp: 1000 magnification.

network structures was reflected in rheological and flotation per- et al., 2013; Meer and Benson, 2007). Ca2+ cations in the interlayer
formance as shown in the previous sections. only allow for crystalline swelling with an interlayer separation of
The Ca2+ cations released by gypsum appear to be present in up to 20 Å. On the other hand, hydration of Na+ cations in the inter-
sufficient quantities to attach to the negative surfaces of the layer can increase that separation to several hundred Angstroms
sodium bentonite particles and change their electrical double (Luckham and Rossi, 1999), which is called osmotic swelling. In
layers. Compression of these layers causes the particles to aggre- addition to this, even small amounts of Ca2+ cations can hold
gate mostly in F–F orientations from original three dimensional together the silicate layers, and build up band-type networks.
honeycomb-like structure in the absence of Ca2+ cations. Ca2+ Large amounts of Ca2+ cations contract these networks forming
cations can also interfere with the interlayer swelling of the compact particle aggregates as illustrated in Fig. 13 (Lagaly and
sodium bentonite, since the exchange of the Na+ cations in the Dékány, 2013b).
interlayer for Ca2+ cations is thermodynamically favourable, and Cryo-SEM images of the flotation froth also show some differ-
complete replacement of Na+ by Ca2+ may happen (Bradshaw ences in bentonite particle aggregation as shown in Figs. 14 and
N. Cruz et al. / Minerals Engineering 77 (2015) 121–130 127

Fig. 12. Cryo-SEM images showing the network structure formed by ore–bentonite (a) and ore–bentonite–gypsum (b) mixtures in flotation pulp: 4000 magnification.

15 at two magnifications. In the slurry with sodium bentonite with many inter-aggregate water sites all over the strings
addition long strings of clay particles were formed from the aggre- (Fig. 17b and 18b). Similar to the ore–bentonite–gypsum mixture,
gation at the borders of the bubbles before breakage during sam- some gel-like strings were observed in the froth sample of ore with
pling, and the addition of gypsum made thinner strings with kaolinite and gypsum addition (Fig. 17b) again suggesting that
more compact aggregates that appeared to be made mostly of E– gypsum took part in the formation of these strings.
E and F–F associations. The concentration or packing of these The presence of these strings was clearer in the froth images
aggregates was also reflected in the mass–water recovery and with a higher magnification (Fig. 18). It seems that in the mixture
recovery by entrainment that showed the transport of more solids containing gypsum, the strings attached to kaolinite agglomerates
into the concentrate (Figs. 1 and 5). The aggregates in ore–ben- increasing the entrainment of clay minerals as observed in the
tonite mixture formed less compacted strings in E–E, F–F and E– water-recovery curves (Fig. 1). This also resulted in a decrease in
F association modes. It is interesting to note that the gel-like copper and gold grade (Figs. 2 and 3). EDS was used to determine
materials were present all over the particles in both the pulp and the element composition of the strings being 0.3 at.% Na, 0.5 at.%
froth samples. They contained 6 at.% Na, 18 at.% Al, 41 at.% Si, Mg, 17 at.% Al, 26 at.% Si, 7 at.% S, 2 at.% K and 47 at.% Ca.
19 at.% Ca and 16 at.% S detected by using EDS. Some appeared as However, the result was not accurate because the spot size of beam
threads between the bentonite strings while some presented like was relatively too large (>2 lm) and the gel-like strings were
separate films on the string surfaces or between the strings. always associated with the kaolinite aggregates making the accu-
However, no gypsum particle has been found in both pulp and rate measurement of individual gel-like strings difficult in our
froth samples. It is assumed that the gypsum reacted with ben- experiment set-up. Apparently, the strings were very high in Ca.
tonite to form the gel-like material. It is speculated that the Ca2+ cations from the gypsum interacted
Figs. 16 and 17 show Cryo-SEM images for the aggregation of with Si, Al, and Na from the clay minerals forming these structures.
kaolinite in the pulp of mixtures with and without gypsum at Similar findings were described by Abdi and Wild (1993); Wild
two magnifications. Fig. 18 shows Cryo-SEM images of froth et al. (1993) who explained that lime and gypsum interacted with
agglomerates from the ore–kaolinite mixture and the ore–kaolin- kaolinite to produce gel structures.
ite–gypsum mixture. Kaolinite aggregates associated in the E–E,
E–F and F–F modes in both pulp (Fig. 16a) and froth (Fig. 18a) of 4. Discussion
the mixture without gypsum with much smaller aggregate parti-
cles as compared to the sample with gypsum addition. The aggre- The two clay minerals used in this work increased the shear
gates in the pulp (Fig. 16b) and froth (Fig. 18b) of the mixture with stress values and viscosities in a similar manner when 15 wt.% ben-
gypsum were much larger than samples without gypsum. The tonite or 30 wt.% kaolinite was mixed with the ore, but there was a
thick long strings were associated mostly in F–F and E–E modes significant difference in the distribution of apparent viscosities at
low shear rates. It seems that this difference in apparent viscosity
was related to the different type of network structures formed by
the clay minerals as shown by the Cryo-SEM images. A lower con-
centration of sodium bentonite particles (15 wt.%) gave higher
shear stresses at low shear rates when compared to a higher con-
centration of kaolinite (30 wt.%) in the slurries. This indicates that
the sodium bentonite aggregates are stronger or more difficult to
break, and create more friction when the slurry starts flowing. At
higher shear rates the mixtures with kaolinite showed slightly lar-
ger shear stress values, and this may be related to the higher
kaolinite concentration in the slurry (30 wt.%). It may be the case
that once the network structures in the mixture with 15 wt.% ben-
tonite were broken, there were less clay mineral particles to create
friction at high shear rates compared with the kaolinite mixture.
The differences between bentonite and kaolinite occur as the result
Fig. 13. Schematic showing the aggregation of the clay mineral layers with
of differences in the clay mineral properties such as particle struc-
increasing attraction: (a) single layers, (b) band-type networks, (c) compact ture, swelling behaviour, surface charge density, and cation
particles (reprinted from Lagaly and Dékány, 2013b). exchange capacity. This study indicates that bentonite aggregates
128 N. Cruz et al. / Minerals Engineering 77 (2015) 121–130

Fig. 14. Cryo-SEM images illustrating the change in froth agglomerates from the ore–bentonite mixture (a) to the ore–bentonite–gypsum mixture (b): 1000 magnification.

Fig. 15. Cryo-SEM images illustrating the change in froth agglomerates from the ore–bentonite mixture (a) to the ore–bentonite–gypsum mixture (b): 4000 magnification.

Fig. 16. Cryo-SEM images showing the kaolinite particle aggregates in the mixtures with ore in the absence (a) and presence of gypsum (b) in flotation pulp: 1000
magnification.

have a greater impact on the hydrodynamics in the flotation cell, of apparent viscosities especially at the low shear rate values since
and are more likely to interfere with the transport of minerals to these values may be present at various locations in the high vol-
the froth. ume of the flotation cell when dealing with non-Newtonian
At pH 10 kaolinite particles did not form the sponge-like struc- slurries.
ture as seen in the bentonite mixture. Kaolinite aggregates were Patra et al. (2010, 2012a and 2012b) investigated the effect of
more isolated compact agglomerates with predominant face–face network structures on flotation by using ores containing fibrous
(F–F) interactions. A higher concentration of these aggregates minerals and systems with nylon fibres. The network structures
(30 wt.%) was needed to match the viscosity values of the network formed in their systems are different to the ones containing clay
structures formed by bentonite at lower concentration (15 wt.%). minerals, but there are some similarities in the way they alter
However, similar viscosities did not translate to similar flotation the hydrodynamics in the flotation cell when sodium bentonite is
performance. In this case it is important to look at the distribution present. These researchers found that network structures
N. Cruz et al. / Minerals Engineering 77 (2015) 121–130 129

Fig. 17. Cryo-SEM images showing the kaolinite particle aggregates in the mixtures with ore in the absence (a) and presence of gypsum (b) in flotation pulp: 8000
magnification.

Fig. 18. Cryo-SEM images showing the change in froth agglomerates from the ore–kaolinite mixture (a) to the ore–kaolinite–gypsum mixture (b): 8000 magnification.

interfered with bubble dispersion, froth formation, and entrap- The detrimental effect of kaolinite was mostly due to the
ment of aggregates. The pores in the networks were small entrainment. The highest entrainment occurred when gypsum
(<20 lm) when compared with the bubble size in the flotation cell was added to the kaolinite mixture with the ore. From the Cryo-
(1–2 mm), meaning that bubbles could not penetrate these pores SEM images aggregates were connected by long strings that
in the network structures. This caused poor bubble dispersion, appeared to be formed by the presence of gypsum. These aggre-
and hindered bubble percolation affecting froth formation (Patra gates did not increase viscosity (Fig. 6), and their entrainment
et al., 2012a). Similar phenomenon may happen with the sponge- resulted in transport of more mass to the froth.
like structure formed by bentonite in Figs. 10a and 11a. The struc- In this study frother and collector additions were maintained
tures were not greater than 20 lm, and could affect bubble disper- the same with and without the addition of clay minerals and there
sion and movement through the pulp. There was little froth formed appeared that the consumption of reagents by clay minerals did
during flotation of the ore containing bentonite, and this could be not have a significant effect on the flotation. This may be associated
attributed to the presence of this ‘‘porous’’ structure. It will be also with the way clay minerals modified the flotation. Kaolinite and
reasonable to assume that some valuable minerals could be locked the mixture of bentonite and gypsum enhanced froth stability
in the cells of the network structure. and the consumption of frother by clay minerals may not affect
The 30 wt.% kaolinite mixture with the ore did not affect froth the frothing in flotation. Bentonite formed a sponge-like structure
formation. The froth was more stable than in the baseline case with a high pulp viscosity interfering with the hydrodynamics in
and there was significant entrainment of particles as confirmed flotation, and more or less frother in the pulp may not affect the
by the mass-recovery data. In this case the overall recovery of gold flotation performance either. Meanwhile, chalcopyrite, the main
and copper was slightly improved (Figs. 3 and 4) due to the copper sulphide mineral in the ore, displays good collectorless
entrainment of particles which at the same time diluted the con- flotation (Zachwieja et al., 1989). The consumption of collector
centrate grade. For this slurry the aggregates did not appear to by clay minerals in this study may not affect the flotation
be interconnected, and the hydrodynamics in the flotation cell significantly.
were not badly affected allowing for entrainment of fine gangue
minerals and the formation of abundant froth.
The addition of gypsum to the bentonite mixture shows that 5. Conclusions
Ca2+ cations inhibited the formation of the sponge-like structure
resulting in lower viscosities values, and froth formation could The presence of bentonite and kaolinite resulted in the forma-
occur. The recovery was improved by both true flotation and tion of different types of aggregates in the mixtures with the ore
entrainment, but grade decreased. This indicates that the type of at pH 10 adjusted with lime. Bentonite aggregates had a porous
network structure is important in flotation and may be manipu- or sponge-like structure with predominant edge–edge (E–E) inter-
lated by addition of electrolytes. actions while kaolinite did not form a particular network structure
130 N. Cruz et al. / Minerals Engineering 77 (2015) 121–130

and its aggregates mostly consisted of face–face (F–F) type associa- Cruz, N., Peng, Y., Farrokhpay, S., Bradshaw, D., 2013. Interactions of clay minerals in
copper–gold flotation: Part 1 – Rheological properties of clay mineral
tions. The structures in the mixture with bentonite had higher vis-
suspensions in the presence of flotation reagents. Miner. Eng. 50–51, 30–37.
cosities across a wider range of low shear stress values which Forbes, E., Davey, K.J., Smith, L., 2014. Decoupling rheology and slime coatings effect
might affect hydrodynamics in certain regions of the flotation cell on the natural flotability of chalcopyrite in a clay-rich flotation pulp. Miner. Eng.
and have a detrimental effect on flotation recovery. Recovery in the 56, 136–144.
Jorjani, E., Barkhordari, H.R., Tayebi Khorami, M., Fazeli, A., 2011. Effects of
mixture of the ore with kaolinite was not affected, but grade aluminosilicate minerals on copper–molybdenum flotation from Sarcheshmeh
decreased due to entrainment. The addition of gypsum to the porphyry ores. Miner. Eng. 24 (8), 754–759.
slurry containing bentonite significantly changed the arrangement Lagaly, G., Dékány, I., 2013a. Colloid clay science. In: Bergaya, F., Theng, B.K.G.,
Lagaly, G. (Eds.), Handbook of Clay Science. Elsevier Science, Amsterdam, The
of aggregates by inhibiting the formation of interconnected net- Netherlands, pp. 243–345 (Chapter 8).
work structures. This led to lower viscosity values with flotation Lagaly, G., Dékány, I., 2013b. Colloid Clay. Science 5, 243–345.
behaviour similar to that of mixtures with kaolinite. There was Luckham, P.F., Rossi, S., 1999. The colloidal and rheological properties of bentonite
suspensions. Adv. Colloid Interface Sci. 82, 43–92.
an improvement in recovery, but grade decreased due to Meer, S.R., Benson, C.H., 2007. Hydraulic conductivity of geosynthetic clay liners
entrainment. exhumed from landfill final covers. J. Geotech. Geoenviron. 133, 550–563.
_
Morris, G.E., Zbik, M.S., 2009. Smectite suspension structural behaviour. Int. J.
Miner. Process. 93 (1), 20–25.
Patra, P., Nagaraj, D.R., Somasundaran, P., 2010. Impact of pulp rheology on selective
Acknowledgments recovery of value minerals from ores. In: Singh, R., Das, A., Banerjee, P.K.,
Bhattachalyya, K.K., Goswani, N.G., (Eds.), Mineral Processing Technology,
The authors gratefully acknowledge the financial support of this Jamshedpur, pp. 1223–1231.
Patra, P., Bhambhani, T., Nagaraj, D.R., Somasundaran, P., 2012a. Impact of pulp
study from the Australian Research Council, Newmont Mining
rheological behavior on selective separation of Ni minerals from fibrous
Corporation and Newcrest Mining Limited as well as the discussion serpentine ores. Colloids Surf. A 411, 24–26.
and suggestion from Dr. Ronel Kappes at Newmont Mining Patra, P., Bhambhani, T., Vasudevan, M., Nagaraj, D.R., Somasundaran, P., 2012b.
Corporation and Dr. David Seaman at Newcrest Mining Limited. Transport of fibrous gangue mineral networks to froth by bubbles in flotation
separation. Int. J. Miner. Process. 104–105, 45–48.
The first author also thanks the scholarship provided by the Plancon, A., Giese, R.F., Snyder, R., 1988. The Hinckley index for kaolinites. Clay
University of Queensland. Miner. 23, 249–260.
Savassi, O.N., Alexander, D.J., Franzidis, J.P., Manlapig, E.V., 1998. An empirical model
for entrainment in industrial flotation plants. Miner. Eng. 11 (3), 243–256.
Schubert, H., 2008. On the optimization of hydrodynamics in fine particle flotation.
References Miner. Eng. 21 (12–14), 930–936.
Shabalala, N.Z.P., Harris, M., Leal Filho, L.S., Deglon, D.A., 2011. Effect of slurry
Abdi, M.R., Wild, S., 1993. Sulphate expansion of lime-stabilized kaolinite: I. rheology on gas dispersion in a pilot-scale mechanical flotation cell. Miner. Eng.
Physical characteristics. Clay Miner. 28, 555–567. 24 (13), 1448–1453.
Alther, G.R., 1986. The effect of the exchangeable cations on the physico-chemical Wang, B., Peng, Y., 2014. The interaction of clay minerals and saline water in coarse
properties of Wyoming bentonites. Appl. Clay Sci. 1, 273–284. coal flotation. Fuel 134, 326–332.
Anderson, C.J., 2008. Flotation in a novel oscillatory baffled column. In: Department Wei, R., Peng, Y., Seaman, D., 2013. The interaction of lignosulfonate dispersants and
of Chemical Engineering. University of Cape Town, Cape Town, p. 249. grinding media in copper–gold flotation from a high clay ore. Miner. Eng. 50–51,
Bakker, C.W., Meyer, C.J., Deglon, D.A., 2009. Numerical modelling of non- 93–98.
Newtonian slurry in a mechanical flotation cell. Miner. Eng. 22 (11), 944–950. Wild, S., Abdi, M.R., Leng-Ward, G., 1993. Sulphate expansion of lime-stabilized
Bakker, C.W., Meyer, C.J., Deglon, D.A., 2010. The development of a cavern model for kaolinite- II. Reaction products and expansion. Clay Miner. 28, 569–583.
mechanical flotation cells. Miner. Eng. 23 (11–13), 968–972. Xu, D., Ametov, I., Grano, S.R., 2011. Detachment of coarse particles from oscillating
Battersby, B.J., Sharp, J.C.W., Webb, R.I., Barnes, G.T., 1994. Vitrification of aqueous bubbles – The effect of particle contact angle, shape and medium viscosity. Int.
suspensions from a controlled environment for electron microscopy: an J. Miner. Process. 101 (1–4), 50–57.
improved plunge-cooling device. J. Microsc.-Oxford 176 (2), 110–120. Xu, N., Fan, R., Smart, R.S.C., 2014. Cryo-SEM investigation of aggregate structures of
Bıçak, Ö., Ekmekçi, Z., Can, M., Öztürk, Y., 2012. The effect of water chemistry on clay minerals in flotation pulp and froth. In: Proc. XXVI International Minerals
froth stability and surface chemistry of the flotation of a Cu–Zn sulfide ore. Int. J. Processing Conference, Ch 2 Flotation Fundamentals, 1–11, Santiago, Chile.
Miner. Process. 102–103, 32–37. Zachwieja, J.B., McCarron, J.J., Walker, G.W., Buckley, A.N., 1989. Correlation
Bradshaw, S.L., Benson, C.H., Scalia, J., 2013. Hydration and cation exchange during between the surface composition and collectorless flotation of chalcopyrite. J.
subgrade hydration and effect on hydraulic conductivity of geosynthetic clay Colloid Interface Sci. 132 (2), 462–468.
liners. J. Geotech. Geoenviron. 139 (4), 526–538. Zhang, M., Peng, Y., 2015. Effect of clay minerals on pulp rheology and the flotation
Connelly, D., 2011. High clay ores: a mineral processing nightmare. Aust. J. Min., of copper and gold minerals. Miner. Eng. 70, 8–13.
28–29

S-ar putea să vă placă și