You are on page 1of 11

Int. J. Miner. Process.

93 (2009) 256–266

Contents lists available at ScienceDirect

International Journal of Mineral Processing


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / i j m i n p r o

A study of mechanisms affecting molybdenite recovery in a bulk copper/


molybdenum flotation circuit
M. Zanin ⁎, I. Ametov, S. Grano, L. Zhou, W. Skinner
Ian Wark Research Institute, University of South Australia (The ARC Special Research Centre for Particle and Material Interfaces), Mawson Lakes Campus, Adelaide, South Australia 5095,
Australia

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

Article history: Molybdenite flotation in the bulk copper/molybdenum flotation circuit at Kennecott Utah Copper was
Received 23 April 2009 studied by means of a combination of plant metallurgical surveys, laboratory flotation tests, mineralogical
Received in revised form 2 October 2009 analysis (QEM-Scan), surface analysis (ToF-SIMS) and contact angle measurements. It was demonstrated
Accepted 3 October 2009
that molybdenite recovery is influenced by flotation feed solids percent and by the mineralogy of the host
rock. Molybdenite recovery was consistently higher at reduced flotation feed solids percent. Furthermore,
Keywords:
the recovery of molybdenite was significantly lower from flotation feeds with high limestone skarn ore
Froth flotation
Molybdenite content. The major factors affecting the flotation recovery of molybdenite from both porphyry and skarn
Sulphide ores copper ores are discussed. It is suggested that the lower flotation recovery of molybdenite compared to the
Ore mineralogy copper sulphide is determined by several factors, including particle morphology, inherent hydrophobicity
and possible formation of slime coatings in the presence of gangue minerals typical of skarn ores.
Implications on plant performance are discussed, and solutions to restore molybdenite recovery presented.
© 2009 Elsevier B.V. All rights reserved.

1. Introduction 1961). The occurrence of large molybdenite crystals in vein controlled


mineralisation as opposed to finely disseminated molybdenite has also
Characteristically, in porphyry copper flotation plants, molybde- been reported (Sutulov, 1975; Podobnik and Shirley, 1982), but no
num exhibits lower recovery than copper, in spite of the apparent clear connection with flotation response has been established. Triffett
natural hydrophobicity of molybdenite (Kelebek, 1988). Furthermore, and Bradshaw (2008) demonstrated that molybdenite particles with
molybdenum recovery displays high variability. While copper high aspect ratio (major axis over minor axis) have higher probability
recovery is usually between 80% and 90%, molybdenum recovery of reporting to the concentrate, and that coarse particles with high
may range between 25% and 85% (Crozier, 1979). Among the copper perimeter to area ratio tend to report to the tailings. It could be noted
sulphide minerals, chalcopyrite usually has higher flotation rate and that aspect ratio itself may not be the critical factor, but it may be a
recovery, but also chalcocite, bornite, digenite and covellite can be proxy for hydrophobicity, as discussed further below.
recovered to values higher than 80% if the relevant electrochemical Ametov et al. (2008) conducted a series of surveys at different
conditions are maintained in the slurry to minimise surface oxidation porphyry copper flotation plants. In the operations investigated, the
(Orwe et al., 1998). Molybdenite recovery, on the contrary, may vary recovery of molybdenite in the rougher/scavengers of the bulk Cu/Mo
significantly from operation to operation, and also within different ore flotation circuit was consistently lower than the recovery of copper
bodies in the same operation. Copper and molybdenum recovery data sulphide, the difference ranging from about 2% to 12% (Ametov et al.,
for a one year period of a typical porphyry copper flotation plant are 2008). Furthermore, a reduction in the flotation feed solids percent
reported in Fig. 1, showing high variability of molybdenum recovery (expressed as solids percent in the slurry by weight) resulted in an
with time. increase in molybdenite recovery, while copper flotation was almost
Extensive research has been carried out in an attempt to link the unaffected (Ametov et al., 2008). The different behaviour of copper
flotation response of molybdenite to the mineral crystal structure, minerals and molybdenite with respect to feed solids percent was
textural features and lithology. The degree of crystallisation is one of explained in terms of flotation hydrodynamics (Ametov et al., 2008).
the factors that have been reported to affect molybdenite flotation
(Hernlund, 1961; Shirley, 1981; Podobnik and Shirley, 1982). Well- 1.1. Mechanisms affecting molybdenite recovery
crystallised molybdenite is considered fast floating, while the almost
amorphous variety is either slow floating or non-floating (Hernlund, The low and highly variable flotation recovery of molybdenum
may be a result of several factors, all related to the properties of the
⁎ Corresponding author. Tel.: +61 8 8302 3263; fax: +61 8 8302 3683. molybdenite (MoS2) mineral. Molybdenite crystal structure consists
E-mail address: massimiliano.zanin@unisa.edu.au (M. Zanin). of hexagonal layers of molybdenum atoms between two layers of

0301-7516/$ – see front matter © 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.minpro.2009.10.001
M. Zanin et al. / Int. J. Miner. Process. 93 (2009) 256–266 257

the particles formed are characterised by strongly hydrophobic and


inert faces and hydrophilic and reactive edges, generated by the
breakage of the covalent bonds. These peculiarities determine the
flotation behaviour of molybdenite, which is a combination of (a)
particle morphology (shape and size) in relation to hydrodynamics
(Ametov et al., 2008), (b) particle inherent hydrophobicity, an
important factor controlled to a degree by the face/edge ratio
(Chander and Fuerstenau, 1972; Hoover, 1980), (c) particle–particle
interactions between molybdenite and gangue minerals (Raghavan
and Hsu, 1984), and (d) particle recovery across the froth phase
(Dippenaar, 1982; Zanin et al., 2008). The possible effect of each
contributing factor is discussed further below, and it is a purpose of
this paper to probe the relative importance of some of these
mechanisms.

Fig. 1. Historical recovery data for a typical porphyry copper plant. Bulk copper/
1.2. Hydrodynamic effects (a)
molybdenum flotation circuit.

In flotation, the rate at which particles are removed from the slurry
by air bubbles can be represented by (Newell and Grano, 2007):
sulphur atoms (Fig. 2). Strong covalent bonds act within S–Mo–S
layers, but only weak van der Waals forces between adjacent S–S dNp
= kNp = −Zpb × Ecoll ð1Þ
sheets (Lince and Frantz, 2000). This strong anisotropy causes dt
preferential cleavage of the molybdenite crystal along the adjacent
S–S sheets. As a result, during grinding, platelet shaped fragments, in which Np is the number of particles in the slurry at time t, k the
exfoliating from larger particles, are generally produced. Furthermore, flotation rate constant, Zpb the collision frequency, and Ecoll the
collection efficiency. The collection efficiency Ecoll can be described as
a product of the collision (Ec), attachment (Ea) and stability (Es)
efficiencies:

Ecoll = Ec × Ea × Es ð2Þ

Hydrodynamic conditions have direct influence on both collision


frequency and collection efficiency. Important hydrodynamic para-
meters are bubble diameter, bubble velocity and turbulent energy
dissipation (Newell and Grano, 2007).
Ametov et al. (2008) argued that, molybdenite particles, due to
their peculiar shape factor, could be more sensitive to hydrodynamic
effects than copper mineral particles. In an agitated slurry, platelet
shaped molybdenite particles may align along streamlines of the
suspending liquid and, therefore, have lower probability of collision
with bubbles. In effect, this hypothesis purports that the hydrody-
namic diameter is the minimum dimension of the platelets. Increasing
turbulence would increase collision frequency and efficiency, and
therefore increase the rate of particle collection. This could be
achieved either by increasing the impeller rotational speed or by
reducing the feed solids percent. Damping of local energy dissipation
occurs in suspensions containing higher volume percent of solids, as
higher volume percent of solids produces higher slurry viscosity,
particularly in the case of interacting particles (Schubert, 1999).
Reducing the feed solids percent reduces the slurry viscosity and
increases turbulence, having in turn a positive effect on particle-
bubble collision efficiency (Ec).

1.3. Inherent hydrophobicity (b)

Due to the cleavage mechanisms of molybdenite, very thin


particles are potentially produced in grinding. The overall degree of
hydrophobicity of molybdenite particles depends on the relative
surface exposure of faces (hydrophobic) and edges (hydrophilic). In
flotation, molybdenite particles with higher exposure of edges will
have lower probability of attachment to air bubbles. On the contrary,
particles with a high face/edge ratio will have higher probability of
being recovered, in agreement with the findings of Triffett and
Bradshaw (2008). The overall degree of hydrophobicity of molybde-
nite particles can also be reduced by the adsorption of metal ions in
Fig. 2. Crystal structure of molybdenite (from Lince and Frantz, 2000). solution (Raghavan and Hsu, 1984). Molybdenite has negative zeta
258 M. Zanin et al. / Int. J. Miner. Process. 93 (2009) 256–266

potential across a wide pH range (Wie and Fuerstenau, 1974). Table 1


However, adsorption of positively charged ions at the edges may Average head grade of quartzite and limestone skarn ore samples.

reduce the magnitude, or even reverse the sign of the zeta potential. Cu Fe Mo S SiO2 Al2O3 CaO K2O MgO
Calcium ions, in particular, have been shown to adsorb in the [%] [%] [ppm] [%] [%] [%] [%] [%] [%]
intermediate to high pH range (Healy, 1984). Adsorbed Ca2+ ions Quartzite 0.50 1.9 600 0.4 63 13.8 2.0 7.0 5.0
on the particle edges may reduce the contact angle and flotability of Skarn 0.45 9.6 21 1.5 44 1.7 24 0.3 2.7
molybdenite particles.
Furthermore, significant deformation of particles may occur in
tumbling mills (Hoover, 1980), and it is common to find bent, flotation feeds (synthetic mineral mixtures and blends of different ore
distorted, or striated molybdenite particles in the flotation feed types) have been carried out at the Ian Wark Research Institute.
(Triffett and Bradshaw, 2008). Under these conditions, higher
exposure of edges may occur, due to the breakage of the covalent
bonds, and the spatial distribution and orientation of the hydrophilic 2. Materials and methods
edges may determine the probability of attachment of particles to air
bubbles on collision. Oily collector is typically added in molybdenite 2.1. Ores
flotation to enhance the mineral's hydrophobicity. The presence of
exposed hydrophilic edges on the bent and distorted particles may The Kennecott Utah Copperton concentrator treats porphyry
also prevent spreading of oil droplets on hydrophobic surfaces, thus copper ore, in which molybdenite occurs as a minor phase. Typically,
reducing the effect of the collector. Mo concentrations range from 0.02% to 0.06%. Distinct zones of
mineralisation and alteration are mined at Kennecott, such that four
1.4. Inter-particle interactions with gangue minerals (c) main ore types have been described (Triffett and Bradshaw, 2008),
including a quartzite ore showing high molybdenum grade
The high reactivity of molybdenite edges may also determine and recovery and a more difficult to float, low grade, limestone
particle–particle interactions with other minerals in the slurry, in the skarn ore.
form of slime coatings. Raghavan and Hsu (1984) showed that, in the In the current study, plant surveys have been undertaken with the
presence of Ca2+ ions, the addition of silica to a system of molybdenite plant processing blends of the four main ore types in different
particles causes a sharp decrease in molybdenite flotation recovery. It proportions, and laboratory flotation tests have been performed both
is possible that the calcium ions act as a ‘bridge’, favouring adhesion on plant slurries and controlled blends of quartzite and skarn ores,
between negatively charged molybdenite and silica particles. This prepared at laboratory scale. Average head grades of the quartzite and
mechanism may be relevant in the flotation of an ore with specific skarn ore samples are reported in Table 1, and the mineralogical
gangue mineralisation and dissolution of ions. Inter-particle interac- composition (modal distribution by QEM-Scan) in Table 2. The two
tions between some of the gangue minerals and molybdenite may ores present distinct gangue minerals: large amounts of quartz and
cause slime coatings on the latter, reducing its flotation recovery. This feldspar in the quartzite ore sample, and actinolite, andradite, talc and
effect could be significant at high pH and high alkalinity, due to the calcite abundant in the limestone skarn ore sample (shown in Table 2
bridging effect of the adsorbed Ca2+ ions. This possibility is explored as the Skarn ore).
in this paper. Single mineral molybdenite samples from WILLYAMA-GEO Dis-
coveries were used to produce coarse molybdenite particles which
1.5. Froth phase recovery (d) were used for contact angle measurement by the sessile drop method.

Inherent hydrophobicity and shape may also play a role in the


2.2. Reagents
transportation of molybdenite particles across the froth phase.
Dippenaar (1982) and Hemmings (1981) found a correlation between
The reagents used in this study, both in the plant and laboratory
size and hydrophobicity of the suspended particles and destabilisation
tests, were: oily collector for molybdenite (generic diesel oil was used
of the froth phase by film rupture. For a given particle size, the greater
in the laboratory tests), at a typical addition rate of 22 g/t, dicresyl-
the contact angle, the greater is the destructive compressive stress
dithiophosphate (S-8989) collector for copper minerals, at a typical
induced on the thin film (Hemmings, 1981). This implies that
addition rate of 20 g/t, and methylisobutyl-carbinol (MIBC) as frother
intermediate to coarse particles having contact angle greater than
(15 to 35 g/t). The pH was adjusted by adding lime, with the exception
90° are effective film breakers in froth flotation (Dippenaar, 1982).
of some diagnostic tests in which KOH was used, as outlined in 2.3.4.
Flat and elongated molybdenite particles may fall into this category,
In the laboratory flotation tests carried out at the Ian Wark Research
and have, for this reason, lower recovery across the froth phase (Zanin
Institute, synthetic process water was used, prepared by dissolving
et al., 2009), also called froth recovery (Savassi et al., 1997). In a
separate investigation (Zanin et al., 2008), it was shown that froth
recovery in the roughers of a porphyry copper flotation plant
Table 2
decreases significantly down the bank (from 60% in the first cells to Mineralogical analysis of quartzite and limestone skarn ore samples.
20% in the last cells), and that froth recovery of molybdenite is
Mineral Formula Modal [%]
generally lower than froth recovery of the copper sulphide. Stability of
the froth phase is therefore a factor which should be taken into Quartzite Skarn
consideration at plant scale, particularly in the rougher/scavengers, Quartz SiO2 24.3 10.9
where major losses of coarse molybdenite occur. In this paper, K_Feldspar KAlSi3O8 38. 7 1.0
however, the focus was on identifying factors which affect molybde- Plagioclase NaxCa(1 − x) (Al,Si)AlSi2O8 8.1 0.02
Montmorillonite (Na,Ca)(Al,Mg)6(Si4O10)3(OH)6 − nH2O 4. 7 0.04
nite recovery in the collection zone, while froth phase issues have
Biotite K(Fe,Mg)3AlSi3O10(F,OH)2 15.0 0.06
been addressed elsewhere (Zanin et al., 2008). Actinolite Ca2(Mg, Fe)5Si8O22(OH)2 0.1 8.2
In the current study, investigation was carried out at Kennecott Andradite (Al) Ca3Al2(SiO4)3 0.14 58.7
Utah Copperton Concentrator. Plant surveys have been undertaken on Talc Mg3Si4O10(OH)2 0.01 0.08
flotation feeds having different mineralogy, and ancillary laboratory Calcite CaCO3 1.0 4.9

flotation tests have been carried out plant slurries. Tests on controlled Modal mineral distribution for the major gangue minerals determined by QEM-Scan.
M. Zanin et al. / Int. J. Miner. Process. 93 (2009) 256–266 259

Table 3 in a cryogenic container until ToF-SIMS analysis was carried out. Since
Composition of synthetic process water compared with a sample of process water from the focus of the study was recovery of coarse molybdenite, the ToF-
the KUCC copper circuit.
SIMS analysis was carried out on the +150 μm particles, which were
Ion Synthetic H2O Process H2O isolated by wet sieving. The surface chemistry of the fast (concentrate
[ppm] [ppm] from cell 1) and slow (concentrate from cell 11) floating particles was
investigated, in order to correlate possible differences in surface
Na+ 1308 1310
Ca++ 798 798 composition to flotation response. It was not possible to analyse
K+ 87 87 molybdenite in the flotation tailings, which ideally contain the non-
Mg++ 129 129 floating molybdenite, due to the extremely low molybdenum grade,
Cl− 1818 1940 i.e. a statistically significant number of molybdenite particles could
SO−−
4 2653 2590
HCO− 189 160
not be analysed. No samples of slurry for ToF-SIMS analysis were
3
TDS 6980 7010 collected during surveys 3–4. Therefore, it was not possible to
pH 7.6 7.2 compare the surface composition of molybdenite in surveys with
Cond. [μ S/cm] 6700 8000 different feed ore blends and gangue mineralogy. This important
aspect of the study was, however, conducted on samples generated at
laboratory scale, as described further below.
salts in demineralised water to reproduce the composition of
Kennecott process water (Table 3). 2.3.2. Laboratory flotation tests on plant slurries
In parallel with the surveys, laboratory flotation tests were
2.3. Methods undertaken on conditioned rougher feed slurries collected in the
plant. The tests were conducted using a 5 l Agitair flotation machine
2.3.1. Plant surveys with forced air supply. The samples were taken from the flotation feed
The Copperton concentrator consists of grinding circuits (SAG box where all reagents except frother were added. An impeller speed
milling followed by ball milling) and two flotation circuits: a bulk of 1000 rpm and air flowrate of 5 l/min were used in the tests. The
copper flotation circuit, where copper/molybdenum concentrate is slurry (conditioned rougher feed) was floated without further
produced, and a molybdenite flotation plant, in which molybdenite in collector addition, and at the same pH (9.5–10) as the plant.
the bulk concentrate is separated from the copper minerals. At the Concentrates were collected after 1, 3, 6 and 10 min. All the flotation
time of investigation, the copper circuit consisted of 5 parallel rows of products were assayed on an unsized and size-by-size basis. Unsized
rougher/scavenger cells. The rougher concentrate was treated in and size-by-size recoveries were calculated.
rougher cleaners (2 parallel rows of cells) without regrinding. The
scavenger concentrate reported to regrinding (ball milling) along 2.3.3. Laboratory flotation tests on reconstructed feeds
with the rougher cleaner tailing. The regrind circuit product was In diagnostic tests carried out at the Ian Wark Research Institute,
upgraded in three scavenger cleaner stages, the final concentrate from samples of quartzite and limestone skarn ores provided by Kennecott
which, combined with the rougher cleaner concentrate, produced the were used to reproduce plant slurries. Three different controlled feed
bulk copper/molybdenum concentrate. The target d80 of the flotation blends were prepared: 100% quartzite, 75:25 quartzite/skarn and
feed was 200 μm, and the pH in the rougher/scavengers was 50:50 quartzite/ skarn ores. The ore blends were crushed using
controlled between 9.5 and 10. A bulk concentrate assaying about laboratory jaw and cone crushers, homogenised and ground in a
25–30% Cu and 2–4% Mo was produced, depending on the ore blend laboratory Galigher tumbling mill, using stainless steel rods as
processed. Since most of the molybdenite losses occurred in the grinding media. Grinding was calibrated to a target d80 = 200 μm, by
rougher/scavengers of the copper circuit, this part of the circuit varying the grind time for each feed type. Synthetic process water
became the focus of investigations. (Table 3) was used to simulate plant conditions. The same pH and
Four plant surveys were undertaken, in which timed lip samples reagents scheme were used as in the tests carried out at Kennecott.
and in-pulp samples were collected from each cell down one of the Flotation tests were carried out as described in 2.3.2.
rougher/scavenger rows. In each survey, four sampling rounds were Similarly to the plant surveys, after each laboratory flotation test
performed, over a period of 90 min, in which multiple samples were the coarse particles (+150 μm) from the first concentrate and the last
collected from the rougher/scavenger tailings to reduce experimental concentrate were collected by wet screening and analysed by ToF-
error. Prior to sampling, plant stability was ensured from the control SIMS. Surface analysis was conducted for the concentrates collected in
room, and plant data (throughput, feed solids percent, air flows, and flotation tests on both 100% quartzite ore and 50:50 blend of quartzite
lip levels) were recorded. All metallurgical samples were initially and skarn ores, to draw a link between molybdenite surface
assayed unsized. Samples from selected surveys were sized by wet/ composition, gangue mineralogy and flotation response .
dry sieving and assayed on a size-by-size basis. Data reconciliation
and mass balance were performed using the software Bilmat 9.3™. 2.3.4. Laboratory flotation tests on coarse molybdenite particles
Surveys 1 and 2 were performed with the plant processing a blend The effect of calcium and magnesium ions in solution and gangue
with high quartzite ore content, while surveys 3 and 4 were carried minerals on the recovery of molybdenite was studied separately.
out with the plant processing a blend with high limestone skarn ore Coarse molybdenite particles (+150 μm) were obtained by dry
content. Surveys 1 and 4 were performed at lower solids percent in screening of a molybdenite concentrate from Kennecott (final product
the flotation feed, while surveys 2 and 3 were carried out at higher
solids percent (Table 4). The percent solids, which under standard Table 4
operating conditions ranged between 30% and 35%, was adjusted by Feed composition and operating conditions during the plant surveys.
varying the water flowrate to the rougher flotation distribution box.
Ore blend Cu Mo Survey Solids pH
The influence of feed solids percent on the recovery of molybdenite
type
from the two different ore blends was studied. [%] [%] # [%]
During survey 1, additional samples of slurry from the concentrate High quartzite 0.3 0.06 1 27 9.8
of cells 1 and 11 were collected for surface analysis by ToF-SIMS. 2 35 10.0
High skarn 0.5 0.06 3 35 10.1
Samples were placed in plastic vials, purged with nitrogen to remove
4 27 10.1
oxygen and frozen in liquid nitrogen. The samples were stored frozen
260 M. Zanin et al. / Int. J. Miner. Process. 93 (2009) 256–266

of the molybdenum plant). Particles were washed in Ensolv (n-propyl skarn ore in the feed blend, but molybdenum recovery decreased
bromide N 93%) and ethanol to clean the surface from any residual more than copper recovery. With a reduction in the feed solids
collector. Samples of quartzite and skarn ore were floated to remove percent, however, it was possible to partially restore molybdenum
copper minerals and molybdenite, and the flotation tailings were recovery to values characteristic of quartzite ore (from 75% to 85%).
collected. The coarse molybdenite particles were then blended into The laboratory tests (Fig. 3c and d) showed generally higher
the tailings of quartzite and skarn ores, to produce synthetic ore recoveries for both copper and molybdenum compared to the plant
containing 0.15% MoS2 by weight. Samples were floated, with the surveys (Fig. 3a and b), but with a similar trend with respect to feed
collectors and frother added according to the conditioning scheme composition and solids percent. Copper recovery was marginally
described in 2.2. Low conductivity water, produced by reverse affected by changes in feed solids percent in the absence of limestone
osmosis, two stages of ion exchange and two stages of activated skarn ore, showing some difference (90% recovery at 35% solids versus
carbon prior to final filtration, was used in these experiments. The 93% recovery at 27% solids) when limestone skarn ore was blended to
water pH was adjusted to 10 by adding KOH, and the concentration of flotation feed. Molybdenum recovery, on the contrary, was always
ions was varied by adding Ca(NO3)2 and Mg(NO3)2. higher at low solids percent. Differences in recovery became
significant (93% versus 85%) in the presence of limestone skarn ore.
2.4. Characterisation techniques Size-by-size analysis of the flotation products was undertaken for
surveys 3 and 4 (feed blend with high skarn ore), in which a
2.4.1. Contact angle of molybdenite particles significant increase in molybdenum recovery was obtained by
Large molybdenite crystals were obtained from WILLYAMA-GEO reducing the feed solids percent. Results (Fig. 4a) revealed increased
Discoveries. Molybdenite crystals were cut and cleaved using a molybdenum recovery across all size ranges, but particularly for the
scalpel, to expose fresh crystal edges and faces. The contact angle was coarse size fractions (+150 μm). The effect on copper was much less
measured at different pH values and Ca2+ concentrations in the pronounced. In the laboratory scale tests, the recovery of molybde-
ranges typically observed in the plant (8–11 for pH and 0–10− 2 M for num was higher than in the plant across all size ranges. Furthermore, a
Ca2+ ions). After conditioning, the water advancing and receding reduction in the feed solids percent produced a significant effect on
contact angle of face and edge of individual molybdenite particles coarse (+150 μm) molybdenum bearing particles, for which recovery
were measured by the sessile drop method. increased from 60% to 80% when the feed solids percent was reduced
from 35% to 27%.
2.4.2. ToF-SIMS Two additional flotation tests were undertaken at laboratory scale,
TOF-SIMS spectra were obtained using a PHI TRIFT II System at 30% and 45% solids in the rougher flotation feed. These concentra-
equipped with a gallium liquid metal ion gun (LMIG) in pulsed mode. tions were achieved by manipulating samples of plant feed collected
For insulating samples the surface charge is compensated using a during survey 3 (35% solids), the former by diluting with process
pulsed electron flood gun. The mineral samples were mounted on to water, the latter by filtering and re-suspending the solids in process
indium foil. The primary beam current employed in the present study water. The trend (Fig. 4d) confirmed what was observed in the
was 600 pA (DC measurement). In static mode, the analysis is previous tests, showing high sensitivity of the coarse molybdenite
confined to the top two monolayers. An excitation voltage of 25 kV particles to feed solids percent. The feed solids percent is an important
was used in un-bunched mode to give a spatial resolution of better driver to molybdenite recovery overall, and particularly in coarse
than 0.5 μm and pulse length adjusted to give a mass resolution (m/ particle size fractions.
Δm) of ~ 4000. Imaging of the sample involved mapping for positive In Fig. 4, bar charts reporting the distribution of copper and
and negative ions for surface regions of mineral particles of interest, molybdenum in the feed are superimposed on the recovery curves.
i.e. MoS2. Several tens of particles in each processing stream are The area of each bar in the charts is proportional to the relative mass
analysed and the data statistically presented in terms of normalised of mineral in the specified particle size range. Considering that, at the
(to total ion yield) intensity of signals and 95% confidence intervals. In standard plant grind, the +150 μm size fraction contains about 20% of
this mode, statistical differences in surface chemistry between the the total molybdenite in the feed, the recovery of this size fraction is
same minerals in different streams or under different pulp conditions critical to the overall molybdenite recovery.
may be discerned. In the present case, we are investigating surface Surface analysis results on the slurry samples collected during
chemistry differences between floating and non(slow)-floating MoS2 survey 1 (in which the flotation feed consisted mainly of quartzite
particles. type ore, and no limestone skarn was present) are reported in Fig. 5.
No significant difference in surface composition between fast
3. Results (concentrate from Cell 1) and slow (concentrate from Cell 11)
floating molybdenites was found in this particular case of the
3.1. Plant Studies and Laboratory Tests on Plant Slurries quartzite ore type. It could therefore be concluded that the smaller
differences in molybdenite recovery noted between plant and
The grade/recovery relationship for copper and molybdenum for laboratory scale (Fig. 3) may be ascribed to changes in hydrodynam-
the four plant surveys and the laboratory flotation tests carried out in ics. However, this conclusion will depend on the feed type, as
parallel with the plant surveys are reported in Fig. 3. The ultimate discussed further below.
recovery (combined rougher/scavenger flotation) for copper and
molybdenum is also reported in Table 5. In the plant surveys, 3.2. Laboratory flotation tests on reconstructed feeds
molybdenum recovery ranged from 75% to 92%, while copper
recovery from 85% to 92%. The recovery of molybdenum was The flotation recoveries of copper and molybdenum in laboratory
consistently lower than copper recovery, the difference increasing at scale tests on blends of quartzite ore and skarn ore are reported in
higher feed solids percent and in the presence of limestone skarn ore. Fig. 6. A sharp decrease in molybdenite recovery was observed at
For both copper and molybdenum, the lowest recovery was achieved increasing concentration of limestone skarn ore in the feed. The
at high solids percent and in the presence of limestone skarn ore ultimate molybdenite recovery decreased from 92% in the absence of
(survey 3). Compared to copper, molybdenum recovery was more skarn ore to 85% with 25% skarn ore and 56% with 50% skarn ore in the
sensitive to the feed solids percent, being significantly enhanced flotation feed. Copper flotation was affected to a much lesser extent.
when solids percent was reduced from 35% to 27%. Both copper and This is in agreement with the plant surveys and tests carried out on
molybdenum recoveries were lower in the presence of limestone plant slurries. It should be noted that, since the limestone skarn ore
M. Zanin et al. / Int. J. Miner. Process. 93 (2009) 256–266 261

Fig. 3. Grade/recovery relationship for copper and molybdenum in the plant surveys (a and b) and in the laboratory tests on plant feed (c and d). The ultimate recovery at the
completion of flotation is also indicated (dashed lines).

contains very little molybdenum (Table 1), it is the molybdenite observations at plant scale, which also showed no surface chemical
particles in the quartzite ore which are depressed in the presence of differences on molybdenite. In the presence of 50% skarn ore, on the
skarn ore. contrary, samples of fast and slow floating molybdenites showed
The ToF-SIMS spectra for the fast floating molybdenite (Con 1) and differences in surface chemistry (Fig. 7). In the fast floating sample
slow floating molybdenite (Con 4) collected in the tests with 0% and (Con 1) there are much more exposed molybdenum and sulphur
50% skarn ore in the flotation feed are reported in Fig. 7. compared to the slow floating sample (Con 4). Furthermore, Ca, K, Fe,
Samples of fast (Con 1) and slow (Con 4) floating molybdenite as well as O and OH groups, were more abundant on the latter sample.
in the presence of quartzite type ore only, showed very small Significantly higher Mg on the surface of molybdenite in the presence
differences in surface chemical composition. The different species of skarn ore was also noted in both fast and slow floating fractions
appear in the same proportion on the surface of the two samples, (Fig. 7a).
within statistical error (Fig. 7). This is in agreement with previous
3.3. Laboratory flotation tests on coarse molybdenite particles
Table 5
Ultimate recovery of copper and molybdenum in plant (after rougher/scavenger The effect of calcium and magnesium ions in solution on coarse
flotation) and laboratory. molybdenite (+150 μm) recovery from quartzite and limestone skarn
ores is presented in Fig. 8. In the case of quartzite ore, the cumulative
Ore blend Survey Solids Plant recovery Lab recovery
type
recovery of molybdenite was high, approaching 90% after 8 min of
# [%] Cu [%] Mo [%] Cu [%] Mo [%] flotation. The addition of calcium and magnesium ions to the pulp did
High quartzite 1 27 92 92 95 95 not have an effect on molybdenite flotation. In contrast, molybdenite
2 35 92 88 94 92 recovery from limestone skarn ore was considerably lower (60%) even
High skarn 3 35 85 75 90 85
in the absence of calcium and magnesium ions, and decreased further
4 27 88 85 93 93
to 30% when Ca2+ and Mg2+ ions were added. Apparently, there is a
262 M. Zanin et al. / Int. J. Miner. Process. 93 (2009) 256–266

Fig. 4. Size-by-size recovery of copper and molybdenum in plant surveys 3 and 4 (after rougher/scavenger flotation) (a and b) and in the laboratory flotation tests on plant feed
(c and d). At laboratory scale, additional feed solids percent values (30% and 45%) were obtained by manipulating plant feed samples. The distribution of copper and molybdenum in
the feed is also reported. Feed blend contained high limestone skarn ore.

synergistic effect between the presence in solution of Ca2+ and Mg2+ particles. Particle size and shape, morphology and surface composi-
ions and gangue minerals of the skarn ore. tion are all contributing factors. Insufficient molybdenite liberation in
The effect of pH and calcium ions on the contact angle of the coarse size fractions was also considered in the first instance as a
molybdenite was also investigated (Fig. 9). The tests showed that possible cause for the low flotation recovery. However, QEM-Scan
faces and edges of molybdenite particles have considerably different analysis of the scavenger tailings (Triffett, private communication)
contact angles. Face contact angle was high (about 100°) and showed that molybdenite in the + 150 μm size fraction mainly
independent of solution pH and calcium concentration. In contrast, consists of liberated particles. This confirms that other mechanisms,
contact angle on the edges was low (maximum 45°), and decreased with as outlined in Section 1.1, play significant roles.
increased pH and calcium ion concentration ([Ca2+] N 4 × 10− 3 M). The hydrodynamic behaviour of the flat molybdenite particles in a
turbulent environment may explain the increase in molybdenum
4. Discussion recovery observed at reduced solids percent (Fig. 3). Higher local
turbulent energy dissipation may increase the collision efficiency for
From an industrial perspective, the most important finding from the flat and elongated particles, as hypothesised in Section 1.2. The
the plant studies was the consistent increase in molybdenite recovery fact that molybdenite recovery in laboratory flotation tests was higher
in rougher/scavenger flotation at reduced feed solids percent. This is than in the plant (Fig. 4) may also be due hydrodynamic effects.
in agreement with previous observations for different flotation plants Compared to the plant cells, laboratory batch flotation cells have
(Ametov et al., 2008). The coarse (+ 150 μm) size fractions were higher local turbulent energy dissipation (Newell and Grano, 2007),
affected the most by changes in feed solids percent, and therefore and therefore higher collision efficiency, Ec, and, according to Eq. (1),
fluctuations in the overall recovery of molybdenite in the plant are higher flotation rate. An interesting feature is that it is the coarse and
determined to a large extent by the flotation response of the coarse liberated molybdenite that is mainly affected, the recovery of which is
M. Zanin et al. / Int. J. Miner. Process. 93 (2009) 256–266 263

Fig. 5. Positive (a) and negative (b) ToF-SIMS normalised intensities for + 150 μm molybdenite particles. Average with 95% confidence, N N 23. Fast floating (Cell 1) and slow floating
(Cell 11) molybdenite from plant survey 1 (no skarn ore in the feed).

Fig. 6. Copper and molybdenum recovery in laboratory flotation tests on reconstructed feeds (blends of quartzite and skarn ores in different ratios). Tests at 35% solids (synthetic
process water used); Agitair 5 l cell; impeller speed 1000 rpm; air 7 l/min. 0% skarn ore = 100% quartzite ore.
264 M. Zanin et al. / Int. J. Miner. Process. 93 (2009) 256–266

Fig. 7. Positive (a) and negative (b) ToF-SIMS normalised intensities for on + 150 μm molybdenite particles. Average with 95% confidence, N N 23. Fast floating (Con 1) and slow
floating (Con 4) coarse particles (N 150 μm) in tests with 100% quartzite ore and 50:50 blend of quartzite and skarn ores.

presumably limited by stability efficiency, Es (Pyke et al., 2003). The scales, suggesting that the different flotation recoveries of molybde-
high surface/bulk ratio of the coarse molybdenite particles may nite at high and low solids percent may be ascribed to hydrodynamic
however be beneficial for the stability of particle–bubble aggregates, effects, possibly arising from the intrinsic shape of the particles, rather
because of a higher surface force/inertial force ratio. Surface forces than from the surface cleaning of slime coatings. Therefore, the
favour attachment of particles to bubbles, while inertial forces hypothesis of slime coatings may hold for some feed types and be less
promote detachment. A high surface/bulk ratio results in high important for others. It is surmised that at low pulp density there is
stability efficiency, even in high turbulent conditions such as in less likelihood of inter-particle interactions causing slime coatings. A
laboratory flotation and at reduced solids percent. reduction in slime coatings on the surface of molybdenite particles
Another outcome of the plant studies was that the effect of a results in increased contact angle. Even a small increase in contact
reduction in feed solids percent on molybdenite recovery was high for angle could cause previously non-floatable coarse particles to become
some feed blends (high limestone skarn) while it was much lower for floatable and the ultimate recovery is therefore increased.
others (high quartzite). This was observed at plant scale (Fig. 3b), in The high concentration of Ca2+ and SO2− 4 in Kennecott process
laboratory flotation tests on plant slurries (Fig. 3d), and in tests on water (Table 3) may also suggest deposition of calcite and gypsum
reconstructed feed blends (Fig. 6). Therefore, it is likely that factors from solution due to oversaturation. This could also be a contributing
other than hydrodynamics and collision efficiency, such as surface factor, but it is not sufficient to justify the different degrees of
modification and particle–particle interactions also play a role. ToF- molybdenite depression observed in the presence of quartzite and
SIMS analysis indicated the presence of elements associated with skarn ores. Diagnostic flotation tests on coarse molybdenite particles
hydrophilic components on the slow floating coarse molybdenite in the presence of limestone skarn ore and different levels of calcium
particles in the presence of skarn ore (Fig. 7). This could be either from and magnesium ions in solution (Fig. 8) corroborate the hypothesis
the deposition of fine slimes or adsorption of dissolved ions. In any that interactions between molybdenite and some gangue minerals in
case, the effect is strongly ore type dependent. In the absence of skarn the skarn ore lead to molybdenite depression. The effect could be
ore, no difference in surface composition between fast and slow induced by metal ions adsorbed at the molybdenite edges, as also
molybdenite was apparent at plant (Fig. 5) and laboratory (Fig. 7) suggested by contact angle measurements carried out separately on
M. Zanin et al. / Int. J. Miner. Process. 93 (2009) 256–266 265

5. Conclusions

The experimental results confirm that the flotation of molybdenite


from porphyry copper ores is more sensitive to the operating
environment than that of the copper sulphide minerals. The recovery
of molybdenum in the rougher/scavengers of the bulk Cu/Mo
flotation circuit at Kennecott Utah Copper is highly variable, and
always lower than the recovery of copper. Particularly low molybde-
num recovery was observed in the presence of limestone skarn ore.
As a general rule, it was found that operating the rougher/scavenger
flotation rows at a lower feed solids percent (27% against 35%)
ensures higher and more stable molybdenite recovery.
Several factors have been identified which could affect molybde-
nite flotation, all of them related to the peculiar properties of
molybdenite. Preferential cleavage along the weakly bound S–S layers
may impart to the molybdenite particles singular hydrodynamic
behaviour (low collision efficiency due to the flat and elongated
particle shape), and anisotropy in hydrophobicity (hydrophobic faces
and hydrophilic, highly reactive, edges). It has been shown that edges
have much lower contact angle compared to faces, in particular at
high pH (pH N 10) and high alkaline conditions ([Ca2+] N 10− 2 M), as
in typical flotation environment.
Fig. 8. Recovery of coarse molybdenite particles (+ 150 μm) in flotation tests on
When limestone skarn ore was blended to the feed in laboratory
synthetic feeds. Molybdenite particles from the final concentrate, after surface cleaning,
were blended with ground quartzite and skarn ore (0.15% MoS2 by weight). Tests were flotation tests, molybdenite recovery decreased significantly, similarly
in demineralised water, at pH 10, in the presence and absence of Ca2+ and Mg2+ ions. to what was observed in the plant. Higher concentrations of Ca, Fe, Mg
and K were measured on the slow floating molybdenite particles
compared to the fast floating, which have been correlated to the
faces and edges (Fig. 9). The edges of molybdenite particles are presence of some gangue minerals typical of the skarn ore. A ‘bridging’
particularly reactive, and adsorption of Ca2+ and/or Mg2+ may have effect of calcium and magnesium adsorbed on the surface of
the combined effect of reducing the overall hydrophobicity of molybdenite and some fine gangue particles in the slurry is possible
molybdenite particles and “bridging” for particle–particle interactions and may lead to the formation of slime coatings, thus reducing the
with the negatively charged gangue minerals. This is also in flotation recovery of molybdenite.
agreement with the findings of Raghavan and Hsu (1984) on In terms of strategies to increase molybdenite recovery at plant
molybdenite/silica mineral systems. Slime coatings are possibly scale, operating the rougher/scavengers at low solid percent was a
causing molybdenite depression in the presence of skarn ore. Minerals solution which gave immediate benefits. This benefit may be a result
which are much more abundant in the limestone skarn ore are: of a combination of hydrodynamic effects (higher local turbulent
actinolite, andradite, talc and calcite (Table 2). These minerals are energy dissipation and collision efficiency), rheology and reduced
potentially responsible for the high Mg, Ca and Fe observed on the formation of slime coatings. It also appears that preventing slimes
surface of molybdenite when floated in the presence of limestone from adsorbing on molybdenite particles during flotation may benefit
skarn ore (Fig. 7). This conclusion would be consistent with the fact molybdenite recovery. This is the subject for future investigation.
that the presence of the above mentioned minerals in the flotation
feed has been correlated to periods of low molybdenite recovery
Acknowledgements
(Triffett et al., 2008).
The authors would like to acknowledge the financial support of the
sponsors of the AMIRA P260E project and the Australian Research
Council (ARC). Support from the Management at Kennecott Utah
Copper and site personnel is also gratefully acknowledged.

References
Ametov, I., Grano, S.R., Zanin, M., Gredelj, S., Magnuson, R., Bolles, T., Triffett, B., 2008. In:
Zuo, W.D., Yao, S.C., Liang, W.F., Cheng, Z.L., Long, H. (Eds.), Copper and
molybdenite recovery in plant and batch laboratory cells in porphyry copper
rougher flotation: Proceedings of XXIV International Mineral Processing Congress,
Beijing, China 24–28 September 2008, volume 1, pp. 1129–1137.
Chander, S., Fuerstenau, D.W., 1972. On the natural flotability of molybdenite. Trans.
Am. Inst. Min. Metall. Engrs. 252, 62–69.
Crozier, R.D., 1979. Flotation reagent practice in primary and by-product molybdenite
recovery. Min. Mag. 140, 174–178.
Dippenaar, A., 1982. The destabilisation of froth by solids. I. The mechanism of film
rupture. Int. J. Miner. Process. 9, 1–14.
Healy, T.W., 1984. Pulp chemistry, surface chemistry, and flotation. Symp. Ser. —
Australas. Inst. Min. Metall. 40, 43–56.
Hemmings, C.E., 1981. On the significance of flotation froth liquid lamella thickness.
Trans. Inst. Min. Metall. Sect. C 90, C96–C102.
Hernlund, R.W., 1961. Extraction of molybdenite from copper flotation products. Q. J.
Colo. Sch. Min. 56, 177–196.
Fig. 9. Contact angle (advancing) of coarse molybdenite particles (+150 μm) measured Hoover, M.R., 1980. Water chemistry effects in the flotation of sulphide ores — a review
in proximity of face and edges, after conditioning at different pH values and Ca2+ and discussion for molybdenite. In: Jones, M.J. (Ed.), Complex Sulphide Ores. IMM,
concentrations. Contact angle measured by means of the sessile drop method. London, pp. 100–112.
266 M. Zanin et al. / Int. J. Miner. Process. 93 (2009) 256–266

Kelebek, S., 1988. Critical surface tension of wetting and of floatability of molybdenite Schubert, H., 1999. On the turbulence-controlled microprocesses in flotation machines.
and sulphur. J. Colloid Interface Sci. 124, 504–514. Int. J. Miner. Process. 56, 257–276.
Lince, J.R., Frantz, P., 2000. Anisotropic oxidation of MoS2 crystallites studied by angle- Shirley, J.F., 1981. Byproduct molybdenite plant design. Can. Min. J. 1981, 27–28 March.
resolved X-ray photoelectron spectroscopy. Tribol. Lett. 9, 211–218. Sutulov, A., 1975. Copper Porphyries. Miller Freeman Publications, San Francisco.
Newell, R., Grano, S., 2007. Hydrodynamics and scale-up in rushton turbine flotation Triffett, B., Bradshaw, D., 2008. The role of morphology and host rock lithology on the
cells: Part 1 — cell hydrodynamics. Int. J. Miner. Process. 81, 224–236. flotation behaviour of molybdenite at Kennecott Utah Copper. AusIMM Publication
Orwe, D., Grano, S.R., Lauder, D.W., 1998. Increasing fine copper recovery at the Ok Tedi Series, 9th International Congress for Applied Mineralogy, ICAM 2008 —
concentrator, Papua New Guinea. Miner. Eng. 11, 171–187. Proceedings, 2008, pp. 465–473.
Podobnik, D.M., Shirley, J.F., 1982. Molybdenite recovery at Cuajone. Min. Eng. Triffett, B., Veloo, C., Adair, B.J.I., Bradshaw, D., 2008. An investigation on the factors
1473–1477 October. affecting the recovery of molybdenite in the Kennecott Utah Copper bulk flotation
Pyke, B., Fornasiero, D., Ralston, J., 2003. Bubble particle heterocoagulation under circuit. Miner. Eng. 21, 832–840.
turbulent conditions. J. Colloid Interface Sci. 265, 141–151. Wie, J.M., Fuerstenau, D.W., 1974. Effect of dextrin on surface properties and the
Raghavan, S., Hsu, L.L., 1984. Factors affecting the flotation recovery of molybdenite flotation of molybdenite. Int. J. Miner. Process. 1, 17–32.
from porphyry copper ore. Int. J. Miner. Process. 12, 145–162. Zanin, M., Gredelj, S., Grano, S.R., 2008. Factors affecting froth stability in mineral flotation
Savassi, O.N., Alexander, D.J., Johnson, N.W., Manlapig, E.V., Franzidis, J.-P., 1997. and implications on minerals recovery: a case study. In: Kuyvenhoven, R., Gomez, C.,
Measurement of froth recovery of attached particles in industrial cells. In: Lauder, Casali, A. (Eds.), Proceedings of Procemin 2008, Santiago (Chile), pp. 197–206.
D. (Ed.), Proceedings of the Sixth Mill Operators Conference, Aust. Inst. Min. Metall. Zanin, M., Wightman, E., Grano, S.R., Franzidis, J.-P., 2009. Quantifying contributions to
Publ., Melbourne, pp. 149–155. froth stability in porphyry copper plants. Int. J. Miner. Process. 91, 19–27.

You might also like