You are on page 1of 13

NANOPARTICLE FLOTATION AIDS FOR PENTLANDITE FINES

Songtao Yang1, Zongfu Dai2, Manqiu Xu2, Robert Pelton1,* and Carla Abarca1
1
Department of Chemical Engineering
McMaster University
1280 Main Street West
Hamilton, ON, L8S4L7
(*Corresponding author: peltonrh@mcmaster.ca)
2
Vale Base Metals Technical Excellence Centre
2060 Flavelle Boulevard, Sheridan Park
Mississauga, ON, L5K1Z9

ABSTRACT

Fine mineral particles are difficult to process by flotation because of their low collision
efficiencies with air bubbles. Decreasing air bubble size and increasing mineral particle size through
flocculation are the most common approaches to improve flotation efficiencies. Herein we demonstrate
that nanoparticle flotation collectors can be engineered to selectively induce fine mineral particle
aggregation and promote flotation.

Flotation tests of clean fine pentlandite (Pn) and Clarabelle Mill (CBM) rock tails and pyrrhotite
(Po) tails that contained significant amount of fine pentlandite have confirmed that the nanoparticles can
serve as both collectors and selective flocculants for fine pentlandite. In the case of clean fine Pn, the
nanoparticle collectors virtually can float almost all Pn fines. Scoping tests on CBM tails have shown
significant improvement in nickel recovery. Addition of nanoparticles as a flotation aid can recover 20% of
the nickel contained in the rock tails and 40% of the Pn contained in the Po tails compared to current mill
operations giving 0% nickel recovery from rock tails and no Pn selectivity from Po tails. Ongoing work
includes attempts to identify optimal strategies for employing nanoparticles in fines flotation.

KEYWORDS

Flotation, nanoparticle collector, pentlandite fines, Clarabelle Mill, rock tails, pyrrhotite tails, nickel
recovery

194
47th Annual Canadian Mineral Processors Operators Conference©, Ottawa, Ontario, January 20-22, 2015

INTRODUCTION

Fine particle flotation is an industry-wide challenge in mineral processing. Fine mineral particles
are difficult to process by flotation because fine particles have low mass and thus low momentum, which
results in their low collision efficiency with air bubbles (Dai, Fornasiero, & Ralston, 1999). In addition,
fine particles have not only high specific surface area but also more active surface area, which causes high
reagent consumption. For base metal sulphide minerals, particle size between 10 µm and 70 µm is a typical
range where high flotation recoveries can be achieved with conventional flotation chemistry and acceptable
cell residence times. For pentlandite, the flotation recoveries decrease sharply when the particle size is
below 10 µm and hydraulic entrainment is the dominant mechanism, giving recoveries proportional to
water recovery and no Pn enrichment. (Jameson, Nguyen, & Ata, 2007)

To improve fine mineral flotation, a number of approaches have been developed to increase the
bubble-particle collision efficiency, including increasing apparent particle size or decreasing bubble size.
The former approaches include selective flocculation, carrier flotation, emulsion flotation, oil extended
flotation and spherical agglomeration. Selective flocculation requires polymeric flocculants to bridge fine
mineral particles and form loose flocs. Selective flocculation-flotation has yet to be widely applied, because
the entrapment of gangue in flocs is too high (Miettinen, Ralston, & Fornasiero, 2010). The use of large
(0.01 to 10 mm) hydrophobic carrier particles that bind and transport fine minerals has been described
including coated polypropylene particles (Valderrama & Rubio, 1998), oil-coal particles (Sen, Ipekoglu, &
Cilingir, 2010) and polymeric beads (Rubio & Hoberg, 1993). These methods have been the topics of
research for many years but have found very limited commercial applications.

In the last few years, the McMaster Interfacial Technologies Group have worked with the Vale
Base Metals Technology Centre to develop hydrophobic, nano-sized (~ 50 nm) polystyrene latex particles
that function as Pn flotation collectors replacing, or in conjunction with, potassium amyl xanthate (PAX).
(Yang, Pelton, Raegen, Montgomery, & Dalnoki-Veress, 2011) By controlling the surface chemistry of the
latex nanoparticles it is possible to achieve selective deposition on Pn surfaces resulting in good flotation
recoveries. PAX is an inexpensive chemical and it is unlikely that more complex nanoparticles offers any
advantage for separation of larger Pn fractions. (Yang, Pelton, et al., 2013) On the other hand, there is
potential for using our nanoparticle flotation collectors in fine particle separations. Historical production
data of Clarabelle Mill (CBM) have shown significant amount of metal losses to tails. Most of the lost Ni
exists in the form of fine liberated Pn. For example, size by size analyses of a set of daily composite
samples show that 43.8% of the Ni lost to the rock tails was in the <38 µm fraction while 80.7% of the Ni
lost to the Po tails was in this size fraction.

In this paper we show that our nanoparticles function as both selective flocculants and as flotation
collectors for fine Pn mineral suspensions. There have been publications describing the use of latex as
flocculants for fine coal particle suspensions. (Attia & Yu, 1991; Laskowski & Castro, 1999; Laskowski &
Yu, 2000; Littlefair & Lowe, 1986) Most of these studies did not involveflotation and the latex surface
chemistries were not designed to give selective deposition in suspensions with mixed particle types.

EXPERIMENTAL
Materials

Chemicals

Styrene (99%, Sigma-Aldrich), n-butyl acrylate and 1-vinylimidazole (VI, ≥ 99%, Sigma -Aldrich)
were purified by vacuum distillation or passing through inhibitor-remover columns. 2, 2'-azobis (2-
methylpropionamidine) dihydrochloride (V50, 97%), cetyltrimethylammonium bromide (CTAB, 95%)
were purchased from Sigma-Aldrich and used as supplied.

Potassium amyl xanthate (PAX); and the frother, UNIFROTHER 250C or FX-160-13, were
donated by Vale Base Metals Technology Development (Mississauga, ON). For flotation tests involving
the CBM tails, Mississauga tap water was used as flotation process water.

195
47th Annual Canadian Mineral Processors Operators Conference©, Ottawa, Ontario, January 20-22, 2015

Samples – Pn fines

Pentlandite (Pn, Ni4.5Fe4.5S8, 27±2% nickel, ≥ 75% in purity) was donated by Vale Base Metals
(Mississauga, ON). The Pn sample was a high-grade concentrate produced in batch flotation of drill core
sample form Pipe deposit. The concentrate was then dry-pulverized to minus 38μm. The area-averaged
mean diameter (D32) of the sample was 6.2 μm as determined by Malvern Mastersizer. The sample was
then cleaned according to a standard procedure (Yang, Pelton, et al., 2013) provided by Vale in order to
remove any residual PAX and potential oxidation products from the particle surfaces.

Samples – Clarabelle Mill (CBM) rock tails and Po tails

CBM Po tails and rock (Rk) tails slurry were taken by personnel of CBM technical support group
and shipped to Sheridan Park for processing. After storing in the lab for several days, the solids have well
settled. The supernatant was decanted, and the wet solids were manually homogenized and then manually
split into test charges containing 250 g dry solids, bagged and stored in a freezer. The different size
fractions of the two tails were separated by Cyclosizer and sieves. Each fraction was chemically measured
by SP-XRF-FUSION.

The particle size distributions of the two tails sample are shown in Figure 1. The overall particle
size distributions of the two samples show that 80% by weight of the Rk tails can pass 199 µm, whereas
80% of the Po tails can pass 49 µm. In rock tails about 45% Pn particles are smaller than 38 µm, and in Po
tails about 80% Pn can pass 38 µm.

100
CBM Rk tails particle size distribution
P80 = 199 µm
80
Percent passing

Overall
60 Pn
Cp
40 Po
Rk
20
A
0
1 10 100 1000
Particle size/ µm
100
CBM Po tails
particle size distribution
80
P80 = 49 µm
Percent passing

Overall
60
Pn
Cp
40
Po
20 Rk
B
0
1 10 100 1000
Particle size/ µm

Figure 1 – Fractionation of CBM Rk tails and Po tails. In rock tails about 45% Pn particle size are
smaller than 38 µm, and in Po tails about 80% Pn can pass 38 µm

196
47th Annual Canadian Mineral Processors Operators Conference©, Ottawa, Ontario, January 20-22, 2015

The average head assay of the two CBM tails samples are summarized in Table 1. Mineralogical
analyses by means of mineral liberation analyzer (MLA) indicate the main valuable minerals are
pentlandite (Pn) and chalcopyrite (Cp), the main gangue minerals are quartz, feldspar, amphibole, biotite
and chlorite, etc., (collectively called rock, Rk), for the Rk tails sample, and the main gangue sulphide
mineral is pyrrhotite (Po) for the Po tails sample.

Table 1 – The average head assay (%) of the two CBM tails samples used in this work
Sample Cu Ni Fe S Cp Pn Po Rk
CBM Rk tails 0.06 0.11 9.82 1.02 0.14 0.23 1.97 97.7
CBM Po tails 0.18 0.80 33.4 17.7 0.46 1.29 43.8 54.4

Nanoparticle Preparation and Characterization

Polystyrene based (i.e., NP203, NP506) and poly(styrene-co-n-butylacrylate) (i.e., NP60 and
NP209) based cationic nanoparticles were prepared by classic emulsion or emulsifier-free polymerization.
(Goodwin, Ottewill, & Pelton, 1979) Poly(styrene-co-vinylimidazole) based (i.e., NP52, NP64 and NP66)
and poly(styrene-co-n-butylacrylate-co-vinylimidazole) based (i.e., NP41 and NP58) nanoparticles were
prepared by emulsion copolymerization or starved-feed semi-batch emulsion copolymerization. The co-
monomer vinylimidazole was designed to functionalize the nanoparticle surface leading to an increased
pentlandite affinity via ligand-Ni complexation. (Yang, Pelton, et al., 2013) The co-monomer n-
butylacrylate was employed to increase the nanoparticle softness. Details of nanoparticle preparation are
available in a thesis. (Yang, 2012)

Nanoparticle diameters (the number in the nanoparticle designation) were determined by dynamic
light scattering (Brookhaven Instruments Corporation, BIC) using a detector angle of 90°. Electrophoretic
mobility measurements were performed by a Zeta PALS instrument (Brookhaven Instruments Corp.) at
25°C in phase analysis light scattering mode. Samples for both dynamic light scattering and electrophoretic
mobility measurements were prepared in clean vials by dispersing approximately 0.25 g/L of nanoparticles
in 5 mM NaCl. According to electrophoretic mobility measurements, all nanoparticles (latex) prepared
were positively charged, reflecting cationic amidine groups from the initiator (V50).

Contact angle measurements (for both sessile water drop methods and attached bubble methods)
were performed on glass microscope slides (Gold Line Microscope Slides, VWR). Glass slides were cut to
approximately 9 mm squares, cleaned and immersed in ~ 500 mg/L nanoparticle suspension in 5 mM NaCl
for 30 min. The treated slides were then immersed in approximately 1 liter of water to remove unbound
nanoparticles. The contact angle measurements were performed using a Krüss contact angle measuring
instrument running Drop Shape Analysis (DSA) 1.80.0.2 software. Advancing water contact angles (θa)
were recorded by forming a water drop on the end of a fine glass capillary tube (the tip diameter ~ 40-50
µm) and pushing the drop to contact with an air dried latex treated glass slide. Attached bubble receding
contact angle (θr) was acquired by forming an air bubble from the capillary and slowly pushing the air
bubble in contact with never-dried treated glass slides that immersed in water, transferring the bubble to the
surface, yielding a receding contact angle.

Contact angle measurements of each nanoparticle collector have suggesting all prepared
nanoparticles are sufficiently hydrophobic for flotation (Yang & Pelton, 2011). For example, NP60
rendered the advancing water contact angle θa of 91° and the attached bubble receding contact angle θr of
68°, suggesting that the nanoparticles were hydrophobic reflecting the nature of poly(styrene-co-
butylacrylate). When VI fraction was added into the nanoparticles, both contact angles were decreased. For
example, NP58 were decreased by about 20°, leading to θa of 72° and θr of 41°. The decrease of contact
angles is possibly due to the hydrophilicity of imidazole groups on nanoparticle surface. With lower ratio of
VI in the composition NP41 gave both intermediate contact angles at θa of 79° and θr of 62°.

197
47th Annual Canadian Mineral Processors Operators Conference©, Ottawa, Ontario, January 20-22, 2015

Fine Pn flocculation studies

Fine Pn flocculation studies were examined by particle size characterization before and after
treating with nanoparticle collectors. The particle size distributions were measured with a Malvern
Mastersizer 2000. 0.2 mL of 38 g/L NP41 was added into 0.5 mL of roughly 0.2 g/mL of washed fine Pn
that dispersed in 650 mL 5 mM Na2CO3 and the particle size was measured after conditioning for 3 minutes.
The control fine Pn without any nanoparticle addition was measured after 30 seconds ultra-sonication.

Flotation of clean fine Pn suspensions

Flotation experiments of fine Pn were conducted in a custom lab-scale flotation apparatus. In a


typical Pn flotation experiment 10 mL of 0.2 g/mL washed fine Pn suspensions and 0.4 mL of nanoparticle
collectors (38 g/L for NP41) were added into 110 mL of 5 mM Na2CO3 in a 150 mL beaker. The mixture
was conditioned for 2.5 minutes to allow the nanoparticles to deposit onto the surface of Pn and flocculate
fine Pn. Following conditioning, 120 µL of 1% UNIFROTH 250C (10 ppm) was added and mixed for an
additional 0.5 minute. Then nitrogen flow was started at a rate of 1.4 L/min through a Corning Pyrex gas
dispersion tube with a 30 mm diameter coarse glass frit attached by a 90-degree elbow. The froth phase was
scraped over the edge of the beaker and captured in a plastic Petri dish. After 1-1.5 minutes, the gas flow
was stopped. The recovered fine Pn were filtered, dried and weighed. Flotation results were expressed as
the recovery, which was calculated from the mass fraction of solids collected in the froth phase.

The dosage of nanoparticles collectors was expressed with a typical unit used in mining industry,
gram per tonne (g/t). We calculate the dosage based on the corresponding mass of ore that would be
required to produce the amount of pure Pn used in the test. In all of the dosage expressions for Pn fines, we
assume the ore contains 1.5 wt.% Pn and the fine Pn particles in our tests are pure pentlandite. For example,
if floating 2 g pure fine Pn, we used 15.2 mg NP41 in dry mass. One tonne of the ore contains 15 kg Pn. To
process 15 kg Pn, the corresponding NP41 used will be 114 g. Thus, the corresponding nanoparticle dosage
was expressed as 114 g/t.

Flotation of CBM rock tails and Po tails

Flotation tests employing CBM Rk tails and Po tails samples were performed in a 1.1 L Denver
flotation cell at the Vale laboratory. A 250 g test sample was charged into the Denver cell topped up with
laboratory tap water to a volume of 1.0 L. The nanoparticle collector was added and conditioned for 4.5
minutes, and then 20 g/t frother FX-160-13 for Rk tails (10 g/t frother for Po tails) was added and
conditioned for an additional 30 seconds. The slurry pH and conductivity were measured after conditioning.
Flotation was initiated by turning on the air at a flowrate of 1.7 L/min for Rk tails (1.5 L/min for Po tails),
and froth paddle (12 rpm) was set at the same time. Scoping flotation test stage involves continuing the
flotation for 12 minutes while maintaining the pulp level throughout the flotation process, so there are only
two products per test: a concentrate and a tails. The more systematic incremental test stage involves
collecting incremental concentrates at the following time intervals: 0.5 min, 1.0 min, 1.5 min, 3.0 min and
6.0 min, with a total flotation time of 12 min, so there are total six products per incremental test: 5
concentrates and a tails. Each product was dried, weighed, chemically measured by SP-XRF-FUSION and
assayed by the mass balance calculation.

A blank test (no collector addition) and a baseline test (PAX alone) were conducted for each tails
sample. This provides the benchmark for assessing the effectiveness of the nanoparticle collector.

RESULTS AND DISCUSSIONS

Pn fine suspension flocculation with nanoparticles

The ability of nanoparticles to induce flocculation of Pn fines was illustrated by comparing the
particle size distributions before and after nanoparticles introduction. The results in Figure 2 showed that

198
47th Annual Canadian Mineral Processors Operators Conference©, Ottawa, Ontario, January 20-22, 2015

D32 increased from 6.2 µm to 21.5 µm after treatment with the nanoparticle collector, approximately 3.5
times. The nanoparticle dose in this experiment corresponded to dry nanoparticle/Pn ratio of 7.6×10-2 wt/wt.
If all the nanoparticles adsorbed on the Pn surfaces, the corresponding fractional surface coverage was
about 1. 3×103 %. This means about 13 monolayers of nanoparticles were formed on the Pn surfaces, if the
nanoparticles were not aggregated. The required dose to get flocculation should be way below this dosage.
Our flotation experiments in the next part were using about dry nanoparticle/Pn ratio of 7.6×10-3 wt/wt, the
corresponding fractional surface coverage was only about 130 %.
10
Mastersizer Fine Pn (-38 µm) Size Distributions
8 Fine Pn + NP41
Volume Percent

Fine Pn D32 = 21.5 µm


6 D32 = 6.2 µm

4
25 oC
2 5 mM Na2CO3

0
0.1 1 10 100 1000

Particle Size/ µm
Figure 2 – Particle size distribution of the fine Pn suspensions before and after addition of nanoparticle
collector NP41 at a dosage of 7.6×10-2 dry latex/dry Pn

Flotation of clean fine Pn suspensions

Figure 3 compares the flotation recoveries of washed fine pentlandite by using a series of
nanoparticle collectors with that by the conventional PAX. Clearly, the addition of nanoparticle collector
facilitated flotation recovery of fine Pn in comparison to the control and PAX. Without collector only
about 28% Pn was recovered by hydraulic entrainment and with PAX the recovery could only increase to
46%, whereas higher recoveries could be achieved in the presence of the nanoparticles, e.g. 83% by NP60
and 93% by NP41.

An obvious observation from Figure 3 is that smaller nanoparticles were more effective for fine Pn.
NP506 (the largest nanoparticle with 506 nm in diameter) is the worst collector in the series, only gave
39% recovery which is slightly higher than the control. NP209 and NP203 gave intermediate recoveries at
about 66% and 54%. The ligand functionalized NP41 gave the best performance, presumably because it is
the smallest nanoparticle and also bears imidazole groups to increase affinity towards Pn. NP58 could not
compete with NP41, although it contains more imidazole moieties. According to the contact angle
measurements NP58 was less hydrophobic than NP41, which explains the advantage of NP41.

One of our previous poly(styrene-co-vinylimidazole) particles NP52 was added for the
comparison. Unlike other soft nanoparticle collectors with the “soft” butylacrylate segments, NP52 is a
hard polystyrene-based plastic particle. The recovery of NP52 for fine Pn was about 54%, which is much
lower than softer ones with similar small size and imidazole moieties, such as NP41 of 93%. This
observation is in agreement with our previous findings (Yang, Razavizadeh, Pelton, & Bruin, 2013) – softer
nanoparticle collectors are superior for one component clean suspensions.

199
47th Annual Canadian Mineral Processors Operators Conference©, Ottawa, Ontario, January 20-22, 2015

100

Percent fine Pn recovered


Flotation 2 g Pn (D32 = 6.2 µm)
130±20 g/t Dosage
Conditioning 3 mins
80 25 °C, 5 mM Na2CO3

60

40

20
Control PAX NP60 NP209 NP41 NP58 NP203 NP506 NP52

Collector type

Figure 3 – Comparison of various nanoparticle collectors with PAX for flotation of fine Pn. Flotation
conditions include: nanoparticles dosage 130 ± 20 g/t, conditioning 3 minutes, 10 mg/L UNIFROTH 250C,
in 5 mM Na2CO3, nitrogen flow rate at 1.4 L/min, and flotation duration 1.0-1.5 minutes. The error bars
were calculated from the replicated tests. The numbers following “NP” are the diameters of the
nanoparticles in nanometer

More detailed comparisons of three smaller nanoparticles with PAX are shown in Figure 4 as a
function of collector dosage. The nanoparticle dosage is an important parameter because it is related to the
collector cost in operations. The same observation as the results shown in Figure 3 was seen – the three
small nanoparticles are more effective collectors than PAX for fine Pn; smallest NP41 with intermediate
imidazole moieties is the best one. The overall trends for all collectors are that the higher the nanoparticle
dosage, the higher the fine Pn recovery. The great improvements in recoveries corresponded to the dosage
increase from about 25 g/t to 100 g/t, whereas only slightly improvements was seen from 100 g/t to 200 g/t.
Note that the two series of experiments in Figure 3 and Figure 4 was performed at a constant and rather
short conditioning time of 3 minutes. The extension of conditioning time may lead to greater nanoparticle
deposition on Pn surface and thus give higher flotation recovery. In turn, the required nanoparticle dosage
for good recoveries may be reduced from current 100 g/t to a lower dose.

100
Percent fine Pn recovered

NP41

80
NP58

60 NP52

PAX
40 Flotation 1.5-2g Pn (D32 = 6.2 µm)
Conditioning 3 mins
Control - no collector pH = 8.5-10

20
0 50 100 150 200
Collector dosage (g/t)

Figure 4 – Fine Pn recovery with PAX, NP52, NP58 and NP41 as a function of collector dosage. Collector
dosage is expressed as grams of dry nanoparticles used for one tonne ore, g/t

200
47th Annual Canadian Mineral Processors Operators Conference©, Ottawa, Ontario, January 20-22, 2015

In an effort to gain insights into the function of nanoparticles for fine Pn, a concentrate sample
produced with NP41 was investigated by using a JEOL JSM-7000F scanning electron microscope (SEM).
As the SEM images shown in Figure 5, the recovered Pn fines indeed form large flocs in the presence of
NP41. The size of the flocs in the SEM observations ranged from tens of micrometers to about 133 µm as
marked in Figure 5A. This size range falls into the size distribution shown in Figure 2 (red line. The NP41
did function as a flocculant for the fine Pn. At a closer look at the images, e.g. Figure 5C and 5D, irregular
roughness is observed on the surface of the formed flocs. The roughness is induced by the deposition of
nanoparticles or nanoparticle aggregates which increases Pn flocs hydrophobicity. Note that NP41 is a “soft”
nanoparticle with a glass transition temperature close to or below room temperature. It is hard to see
individual spherical particles after air drying at the room temperature.

44 µm
133 µm

84 µm 59 µm
101 µm

A B

C D

Figure 5 – SEM images of dried samples collected from recovered pentlandite fines using NP41 as
the collector. The NP41 functions as both flocculant and collector. The sizes of formed flocs are marked in
the images

In summary, this set of experiments illustrate that the poly(styrene-co-n-butylacrylate-co-


vinylimidazole) nanoparticles or poly(styrene-co-vinylimidazole) nanoparticles could serve as both
flocculants and effective flotation collectors for fine Pn. Both types of nanoparticles perform better than
conventional collector PAX. To now, our results have involved suspensions of clean fine Pn particles.
Presented below are results involving the flotation of the slurry containing fine Pn from a natural nickel
sulphide ore concentrator. The challenge is to obtain good recoveries with a high nickel grade.

201
47th Annual Canadian Mineral Processors Operators Conference©, Ottawa, Ontario, January 20-22, 2015

Flotation of CBM Rk tails

Scoping flotation tests

Figure 6 shows the results of selected scoping flotation tests with NP52, PAX, NP52+PAX,
NP66+PAX by varying collector dosages for CBM Rk tails. PAX alone did not float anything. In contrast,
with the addition of any nanoparticles the nickel recovery can reach up to 21%. The performance of
nanoparticles alone (NP52) was not very well. For example, with a relatively low dosage of 25 g/t NP52,
the nickel recovery was only about 7%. With the combination of nanoparticles and PAX, the nickel
recoveries were increased, ranging from 17% to 21%. However, in NP52-PAX combination, when the
NP52 dosage was increased from 25 g/t, through 50 g/t and 100 g/t to 250 g/t, we did not see much increase
in the nickel recovery. Even the lowest NP dosage, “NP52 (25) + PAX (40)”, could give up to 20% nickel
recovery. This recovery is almost identical to that achieved with 10 times nanoparticle dosage “NP52 (250)
+ PAX (40)”. This means small amount of nanoparticles in combination of PAX can notably improve
nickel recoveries for the Rk tails. NP66a and NP66b are similar nanoparticle collectors with few
alternations from the preparation recipe of NP52. Two sets of the scoping tests at 100 g/t nanoparticles +
PAX (40) were performed. The nickel recoveries were consistent with using NP52 + PAX. This means the
nanoparticle preparations and the flotation performances are repeatable.

25
CBM Rk tails scoping flotation tests
20
Ni recovery (%)

PAX (40) alone


NP52 (25) alone
15 NP52 (25) + PAX (40)
NP52 (50) + PAX (40)
10 NP52 (100) + PAX (40)
NP52 (250) + PAX (40)
5 NP66a (100) + PAX (40)
NP66b (100) +PAX (40)
0
0 5 10 15 20 25
Mass pull (%)
Figure 6 – Comparison of NP52, PAX, NP52+PAX, NP66+PAX by varying collector dosages for scoping
flotation tests of CBM Rk tails. Collector dosage is indicated in parenthesis as grams of dry nanoparticles
per tonne tailings, g/t

Incremental flotation tests

Figure 7 shows the incremental flotation results of of using NP52+PAX and NP58+PAX for CBM
Rk tails. Two data points from the corresponding scoping tests with 500 g/t NP without PAX were added in
the figure. These two points fall on the curves. Again, the combination with NPs and PAX gave slightly
higher nickel recoveries. This means that using higher NP dosage without PAX and using lower NP
dosages with PAX can achieve similar nickel recoveries. Of course using some PAX together with lower
NP dosage is much cheaper than using NP alone at high dosage.

202
47th Annual Canadian Mineral Processors Operators Conference©, Ottawa, Ontario, January 20-22, 2015

25
CBM Rk tails incremental flotation tests
20

Ni recovery (%) 15

NP52 (500)
10
NP52 (50) + PAX (40)
NP52 (250) + PAX (40)
5 NP58 (500)
NP58 (100) + PAX (40)
0
0 5 10 15 20 25
Mass pull (%)
Figure 7 – The results of using NP52+PAX and NP58+PAX for incremental flotation tests of CBM Rk
tails. The two non-filled points were the results of the scoping tests for comparison. Collector dosage is
expressed in parenthesis as grams of dry nanoparticles per tonne tailings, g/t

Flotation of CBM Po tails

Scoping flotation tests

Figure 8 shows the scoping flotation results of using NP58, NP62, NP66, NP64, PAX and
corresponding NP+PAX combinations as collectors for CBM Po tails. NP58 and NP62 are soft
poly(styrene-co-n-butylacrylate-co-vinylimidazole) based nanoparticles, whereas NP66 and NP64 are hard
poly(styrene-co-vinylimidazole) based nanoparticles. The control (None) without any collector gave a low
recovery, suggesting some materials in these Po tails are hydrophobic and can float without any collector
addition. The PAX alone gave an intermediate recovery, suggesting with extended flotation time with PAX,
some lost Ni in thee Po tails is still able to be recovered. Note these CBM Po tails are much finer than CBM
Rk tails. Soft NP58 and NP62 show more collecting power than hard NP66 and NP64 for all fine materials
(including Pn, Cp and Po) in the flotation.

Figure 8 is a plot of Pn recovery versus Po recovery, which shows Pn versus Po selectivity. In the
figure all the hollow symbols are the results of using NPs alone at a 500 g/t dosage. NPs alone have better
selectivity than the use of NPs together with PAX (filled symbols), although the Pn recoveries are slightly
lower. In all cases, “NP64 (250) + PAX (15)” gave best selectivity and acceptable Pn recovery, about 37%
Pn and only 4.7% Po recovered. NP66 (500) alone also gave good Pn recovery and acceptable selectivity,
about 41% Pn and 10% Po recovered.

203
47th Annual Canadian Mineral Processors Operators Conference©, Ottawa, Ontario, January 20-22, 2015

60

Pn recovery (%)
40

20

CBM Po tails scoping flotation tests


0
0 10 20 30 40 50 60
Po recovery (%)
None PAX (30)
NP58 (500) NP58 (250) + PAX (15)
NP62 (500) NP62 (250) + PAX (15)
NP66 (500) NP66 (250) + PAX (15)
NP64 (500) NP64 (250) + PAX (15)
Figure 8 – Comparison of NP58, NP62, NP66 and NP64 with PAX and corresponding PAX and
nanoparticle combinations for scoping flotation tests of CBM Po tails. Nanoparticles alone gave better Pn
selectivity against Po than the cases using PAX or NP-PAX combinations

Incremental flotation tests

More detailed information of using NP66 as flotation aids for CBM Po tails were given in Figure 9
by incremental flotation tests. This set of tests was performed at the same flotation condition except varying
NP66+PAX dosages. A high NP66 dosage of 500 g/t without PAX gave good selectivity (40% Pn recovery
at 10% Po recovery). Combinations of NP66 with PAX did not improve the performance compared with
using PAX alone.
90
NP66 (500)
PAX (30) alone
Pn recovery (%)

NP66 (50) + PAX (10)


60
NP66 (100) + PAX (30)

30

CBM Po tails
Incremental flotation tests
0
0 30 60 90
Po recovery (%)

Figure 9 – The results of using NP66, PAX and NP66+PAX by varying collector dosages for
incremental flotation tests of CBM Po tails. Collector dosage is expressed in parenthesis as grams of
dry nanoparticles per tonne tailings, g/t

204
47th Annual Canadian Mineral Processors Operators Conference©, Ottawa, Ontario, January 20-22, 2015

CONCLUDING REMARKS

This work has extended our initial studies with model suspensions (glass beads, clean pentlandite)
to more realistic CBM tails that contain pentlandite fines. The main conclusions drawn from this work
include:
1. Nanoparticles could function as both selective flocculants and collectors for fine pentlandite.
2. Nanoparticle collectors with 25-50 g/t dosage gave up to 20% nickel recoveries for CBM rock
tails whereas PAX was ineffective – our most encouraging result.
3. Nanoparticles dosage of 500 g/t gave about 40% Pn recoveries for CBM Po tails, mostly present
as pentlandite.

Ongoing work includes batch flotation tests using full circuit simulation flow sheet for natural
CBM sulfide ore and mini-plant evaluation of our best nanoparticles to identify optimal nanoparticle
dosage to make this technology cost effective.

ACKNOWLEDGEMENTS

We thank the Collaborative Research and Development (CRD) Grants from Natural Sciences and
Engineering Research Council of Canada (NSERC) in collaboration with VALE Base Metals for funding.
We also thank Mr. Andy Lee for mineralogy analysis of the Clarabelle Mill rock tails and pyrrhotite tails.

REFERENCES

Attia, Y., & Yu, S. I. (1991). Adsorption Thermodynamics of A Hydrophobic Polymeric Flocculant on
Hydrophobic Colloidal Coal Particles. [Article]. Langmuir, 7(10), 2203-2207. doi:
10.1021/la00058a038

Dai, Z. F., Fornasiero, D., & Ralston, J. (1999). Particle-bubble attachment in mineral flotation. [Article].
Journal of Colloid and Interface Science, 217(1), 70-76. doi: 10.1006/jcis.1999.6319

Goodwin, J. W., Ottewill, R. H., & Pelton, R. (1979). Studies on the Preparation and Characterization of
Monodisperse Polystyrene Lattices .5. Preparation of Cationic Lattices. Colloid and Polymer
Science, 257(1), 61-69.

Jameson, G. J., Nguyen, A. V., & Ata, S. (2007). The Flotation of Fine and Coarse Particles, Froth
Flotation: A Century of Innovation. Littleton, Colorado, USA: Society for Mining, Metallurgy,
and Exploration.

Laskowski, J. S., & Castro, S. H. (1999). Aggregation of inherently hydrophobic solids using hydrophobic
agglomerants and flocculants. Montreal: Canadian Inst Mining, Metallurgy and Petroleum.

Laskowski, J. S., & Yu, Z. M. (2000). Oil agglomeration and its effect on beneficiation and filtration of
low-rank/oxidized coals. [Article]. International Journal of Mineral Processing, 58(1-4), 237-252.
doi: 10.1016/s0301-7516(99)90040-6

Littlefair, M. J., & Lowe, N. R. S. (1986). On the selective flocculation of coal using polystyrene latex.
[Article]. International Journal of Mineral Processing, 17(3-4), 187-203. doi: 10.1016/0301-
7516(86)90056-6

205
47th Annual Canadian Mineral Processors Operators Conference©, Ottawa, Ontario, January 20-22, 2015

Miettinen, T., Ralston, J., & Fornasiero, D. (2010). The limits of fine particle flotation. Minerals
Engineering, 23(5), 420-437.

Rubio, J., & Hoberg, H. (1993). The process of separation of fine mineral particles by flotation with
hydrophobic polymeric carrier. International Journal of Mineral Processing, 37(1), 109-122.

Sen, S., Ipekoglu, U., & Cilingir, Y. (2010). Flotation of Fine Gold Particles by the Assistance of Coal-Oil
Agglomerates. Separation Science and Technology, 45(5), 610-618.

Valderrama, L., & Rubio, J. (1998). High intensity conditioning and the carrier flotation of gold fine
particles. International Journal of Mineral Processing, 52(4), 273-285.

Yang, S. (2012). Nanoparticle Flotation Collectors. Ph.D., McMaster Unviersity, Hamilton, Ontario.

Yang, S., Pelton, R., Abarca, C., Dai, Z. F., Montgomery, M., Xu, M. Q., & Bos, J. A. (2013). Towards
nanoparticle flotation collectors for pentlandite separation. [Article]. International Journal of
Mineral Processing, 123, 137-144. doi: 10.1016/j.minpro.2013.05.007

Yang, S., Pelton, R., Raegen, A., Montgomery, M., & Dalnoki-Veress, K. (2011). Nanoparticle Flotation
Collectors: Mechanisms Behind a New Technology. [Article]. Langmuir, 27(17), 10438-10446.
doi: 10.1021/la2016534

Yang, S., & Pelton, R. H. (2011). Nanoparticle Flotation Collectors II: The Role of Nanoparticle
Hydrophobicity. Langmuir, 27(18), 11409-11415.

Yang, S., Razavizadeh, B. B. M., Pelton, R., & Bruin, G. (2013). Nanoparticle Flotation Collectors-The
Influence of Particle Softness. [Article]. Acs Applied Materials & Interfaces, 5(11), 4836-4842.
doi: 10.1021/am4008825

206

You might also like