You are on page 1of 10

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 in fo abstract

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,
Received in revised form 2 October mineralogical analysis (QEM-Scan), surface analysis (ToF-SIMS) and contact angle measurements. It
2009 Accepted 3 October 2009 was demonstrated 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
Keywords: feed solids percent. Furthermore, the recovery of molybdenite was significantly lower from flotation
Froth flotation
feeds with high limestone skarn ore content. The major factors affecting the flotation recovery of
Molybdenite
Sulphide ores
molybdenite from both porphyry and skarn copper ores are discussed. It is suggested that the lower
Ore flotation recovery of molybdenite compared to the copper sulphide is determined by several factors,
mineralogy 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.

controlled mineralisation as opposed to finely disseminated


1. Introduction
molybdenite has also been reported (Sutulov, 1975; Podobnik and
Shirley, 1982), but no clear connection with flotation response
Characteristically, in porphyry copper flotation plants,
has been established. Triffett and Bradshaw (2008) demonstrated
molybde- num exhibits lower recovery than copper, in spite of the
that molybdenite particles with high aspect ratio (major axis over
apparent natural hydrophobicity of molybdenite (Kelebek, 1988).
minor axis) have higher probability of reporting to the
Furthermore, molybdenum recovery displays high variability.
concentrate, and that coarse particles with high perimeter to area
While copper recovery is usually between 80% and 90%,
ratio tend to report to the tailings. It could be noted that aspect
molybdenum recovery may range between 25% and 85%
ratio itself may not be the critical factor, but it may be a proxy for
(Crozier, 1979). Among the copper sulphide minerals,
hydrophobicity, as discussed further below.
chalcopyrite usually has higher flotation rate and recovery, but
Ametov et al. (2008) conducted a series of surveys at different
also chalcocite, bornite, digenite and covellite can be recovered
porphyry copper flotation plants. In the operations investigated,
to values higher than 80% if the relevant electrochemical
the recovery of molybdenite in the rougher/scavengers of the
conditions are maintained in the slurry to minimise surface
bulk Cu/Mo flotation circuit was consistently lower than the
oxidation (Orwe et al., 1998). Molybdenite recovery, on the
recovery of copper sulphide, the difference ranging from about
contrary, may vary significantly from operation to operation, and
2% to 12% (Ametov et al., 2008). Furthermore, a reduction in the
also within different ore bodies in the same operation. Copper
flotation feed solids percent (expressed as solids percent in the
and molybdenum recovery data for a one year period of a typical
slurry by weight) resulted in an increase in molybdenite recovery,
porphyry copper flotation plant are reported in Fig. 1, showing
while copper flotation was almost unaffected (Ametov et al.,
high variability of molybdenum recovery with time.
2008). The different behaviour of copper minerals and
Extensive research has been carried out in an attempt to
molybdenite with respect to feed solids percent was explained in
link the flotation response of molybdenite to the mineral
terms of flotation hydrodynamics (Ametov et al., 2008).
crystal structure, textural features and lithology. The degree of
crystallisation is one of the factors that have been reported to
affect molybdenite flotation (Hernlund, 1961; Shirley, 1981; 1.1. Mechanisms affecting molybdenite recovery
Podobnik and Shirley, 1982). Well- crystallised molybdenite is
considered fast floating, while the almost amorphous variety is The low and highly variable flotation recovery of
either slow floating or non-floating (Hernlund, molybdenum may be a result of several factors, all related to
the properties of the molybdenite (MoS 2) mineral. Molybdenite
crystal structure consists of hexagonal layers of molybdenum
1961). The occurrence of large molybdenite crystals in vein atoms between two layers of
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
sulphur atoms (Fig. 2). Strong covalent bonds act within S– Grano, 2007):
Mo–S

layers, but only weak van der Waals forces between


adjacent S–S
sheets (Lince and Frantz, 2000). This strong anisotropy causes
S–S sheets. As a result, during grinding, platelet shaped in which Np is the number of particles in the slurry at time t, k
fragments, exfoliating from larger particles, are generally the
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 solution (Raghavan and Hsu,
1984). Molybdenite has negative zeta
Fig. 2. Crystal structure of molybdenite (from Lince and Fra00).
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 Average head grade of quartzite and limestone skarn ore samples.

may reduce the magnitude, or even reverse the sign of the Cu Fe Mo S SiO Al2O3 CaO K2O MgO
zeta potential. Calcium ions, in particular, have been shown to [%] [%] [ppm] [%] 2 [%] [%] [%] [%]
adsorb in the intermediate to high pH range (Healy, 1984). [%]
Adsorbed Ca2+ ions on the particle edges may reduce the Quartzite 0.50 1.9 600 0.4 63 13.8 2.0 7.0 5.0
contact angle and flotability of molybdenite particles. Skarn 0.45 9.6 21 1.5 44 1.7 24 0.3 2.7
Furthermore, significant deformation of particles may occur in
tumbling mills (Hoover, 1980), and it is common to find bent,
distorted, or striated molybdenite particles in the flotation feed flotation feeds (synthetic mineral mixtures and blends of
(Triffett and Bradshaw, 2008). Under these conditions, higher different ore types) have been carried out at the Ian Wark
exposure of edges may occur, due to the breakage of the Research Institute.
covalent bonds, and the spatial distribution and orientation of the
hydrophilic edges may determine the probability of attachment of
particles to air bubbles on collision. Oily collector is typically 2. Materials and Thethods
added in molybdenite flotation to enhance the mineral's
hydrophobicity. The presence of exposed hydrophilic edges on 2.1. Ores
the bent and distorted particles may also prevent spreading of oil
droplets on hydrophobic surfaces, thus reducing the effect of the The Kennecott Utah Copperton concentrator treats
collector. porphyry copper ore, in which molybdenite occurs as a minor
phase. Typically, Mo concentrations range from 0.02% to
1.4. Inter-particle interactions with gangue minerals (c) 0.06%. Distinct zones of mineralisation and alteration are
mined at Kennecott, such that four main ore types have been
The high reactivity of molybdenite edges may also described (Triffett and Bradshaw, 2008), including a quartzite
determine particle–particle interactions with other minerals in ore showing high molybdenum grade and recovery and a
the slurry, in the form of slime coatings. Raghavan and Hsu more difficult to float, low grade, limestone skarn ore.
(1984) showed that, in the presence of Ca2+ ions, the addition of In the current study, plant surveys have been undertaken with
silica to a system of molybdenite particles causes a sharp the plant processing blends of the four main ore types in different
decrease in molybdenite flotation recovery. It is possible that proportions, and laboratory flotation tests have been performed
the calcium ions act as a ‘bridge’, favouring adhesion between both on plant slurries and controlled blends of quartzite and
negatively charged molybdenite and silica particles. This skarn ores, prepared at laboratory scale. Average head grades of
mechanism may be relevant in the flotation of an ore with the quartzite and skarn ore samples are reported in Table 1, and
specific gangue mineralisation and dissolution of ions. Inter- the mineralogical composition (modal distribution by QEM-Scan)
particle interac- tions between some of the gangue minerals in Table 2. The two ores present distinct gangue minerals: large
and molybdenite may cause slime coatings on the latter, amounts of quartz and feldspar in the quartzite ore sample, and
reducing its flotation recovery. This effect could be significant actinolite, andradite, talc and calcite abundant in the limestone
at high pH and high alkalinity, due to the bridging effect of the skarn ore sample (shown in Table 2 as the Skarn ore).
adsorbed Ca2+ ions. This possibility is explored in this paper. Single mineral molybdenite samples from WILLYAMA-GEO
Dis- coveries were used to produce coarse molybdenite
1.5. Froth phase recovery (d) particles which were used for contact angle measurement by
the sessile drop method.
Inherent hydrophobicity and shape may also play a role in the
transportation of molybdenite particles across the froth phase.
2.2. Reagents
Dippenaar (1982) and Hemmings (1981) found a correlation
between size and hydrophobicity of the suspended particles and
The reagents used in this study, both in the plant and
destabilisation of the froth phase by film rupture. For a given
laboratory tests, were: oily collector for molybdenite (generic
particle size, the greater the contact angle, the greater is the
diesel oil was used in the laboratory tests), at a typical addition
destructive compressive stress induced on the thin film
rate of 22 g/t, dicresyl- dithiophosphate (S-8989) collector for
(Hemmings, 1981). This implies that intermediate to coarse
copper minerals, at a typical addition rate of 20 g/t, and
particles having contact angle greater than 90° are effective
methylisobutyl-carbinol (MIBC) as frother (15 to 35 g/t). The pH
film breakers in froth flotation (Dippenaar, 1982). Flat and
was adjusted by adding lime, with the exception of some
elongated molybdenite particles may fall into this category, and
diagnostic tests in which KOH was used, as outlined in 2.3.4. In
have, for this reason, lower recovery across the froth phase
the laboratory flotation tests carried out at the Ian Wark
(Zanin et al., 2009), also called froth recovery (Savassi et al.,
Research Institute, synthetic process water was used,
1997). In a separate investigation (Zanin et al., 2008), it was shown
prepared by dissolving
that froth recovery in the roughers of a porphyry copper flotation
plant decreases significantly down the bank (from 60% in the first
cells to 20% in the last cells), and that froth recovery of
Table 2
molybdenite is generally lower than froth recovery of the copper Mineralogical analysis of quartzite and limestone skarn ore samples.
sulphide. Stability of the froth phase is therefore a factor which
should be taken into consideration at plant scale, particularly in Mineral Formula Modal [%]

the rougher/scavengers, where major losses of coarse Quartzite Skarn


molybdenite occur. In this paper, however, the focus was on
Quartz SiO2 24.3 10.9
identifying factors which affect molybde- nite recovery in the K_Feldspar KAlSi3O8 38. 7 1.0
collection zone, while froth phase issues have been addressed Plagioclase NaxCa(1 − x) (Al,Si)AlSi2O8 8.1 0.02
elsewhere (Zanin et al., 2008). Montmorillonite (Na,Ca)(Al,Mg)6 (Si4 O10)3 (OH)6 −nH2 O 4. 7 0.04
In the current study, investigation was carried out at Biotite K(Fe,Mg)3AlSi3 O10(F,OH)2 15.0 0.06
Kennecott Utah Copperton Concentrator. Plant surveys have Actinolite Ca2(Mg, Fe)5 Si8 O22(OH)2 0.1 8.2
been undertaken on flotation feeds having different mineralogy, Andradite (Al) Ca3Al2(SiO4)3 0.14 58.7
Talc Mg3Si4O10(OH)2 0.01 0.08
and ancillary laboratory flotation tests have been carried out plant
Calcite CaCO3 1.0 4.9
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.


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

of the molybdenum plant). Particles were washed in Ensolv (n- skarn ore in the feed blend, but molybdenum recovery
propyl bromide N 93%) and ethanol to clean the surface from decreased more than copper recovery. With a reduction in the
any residual collector. Samples of quartzite and skarn ore were feed solids percent, however, it was possible to partially
floated to remove copper minerals and molybdenite, and the restore molybdenum recovery to values characteristic of
flotation tailings were collected. The coarse molybdenite quartzite ore (from 75% to 85%).
particles were then blended into the tailings of quartzite and The laboratory tests (Fig. 3c and d) showed generally
skarn ores, to produce synthetic ore containing 0.15% MoS 2 higher recoveries for both copper and molybdenum compared
by weight. Samples were floated, with the collectors and to the plant surveys (Fig. 3a and b), but with a similar trend
frother added according to the conditioning scheme described with respect to feed composition and solids percent. Copper
in 2.2. Low conductivity water, produced by reverse osmosis, recovery was marginally affected by changes in feed solids
two stages of ion exchange and two stages of activated percent in the absence of limestone skarn ore, showing some
carbon prior to final filtration, was used in these experiments. difference (90% recovery at 35% solids versus 93% recovery at
The water pH was adjusted to 10 by adding KOH, and the 27% solids) when limestone skarn ore was blended to flotation
concentration of ions was varied by adding Ca(NO 3)2 and feed. Molybdenum recovery, on the contrary, was always
Mg(NO3)2. higher at low solids percent. Differences in recovery became
significant (93% versus 85%) in the presence of limestone
2.4. Characterisation techniques skarn ore.
Size-by-size analysis of the flotation products was
2.4.1. Contact angle of molybdenite particles undertaken for surveys 3 and 4 (feed blend with high skarn
Large molybdenite crystals were obtained from WILLYAMA- ore), in which a significant increase in molybdenum recovery
GEO Discoveries. Molybdenite crystals were cut and cleaved was obtained by reducing the feed solids percent. Results (Fig.
using a scalpel, to expose fresh crystal edges and faces. The 4a) revealed increased molybdenum recovery across all size
contact angle was measured at different pH values and Ca2+ ranges, but particularly for the coarse size fractions (+150 μm).
concentrations in the ranges typically observed in the plant The effect on copper was much less pronounced. In the
(8–11 for pH and 0–10−2M for Ca2+ ions). After conditioning, laboratory scale tests, the recovery of molybde- num was
the water advancing and receding contact angle of face and higher than in the plant across all size ranges. Furthermore, a
edge of individual molybdenite particles were measured by the reduction in the feed solids percent produced a significant
sessile drop method. effect on coarse (+150 μm) molybdenum bearing particles, for
which recovery increased from 60% to 80% when the feed solids
2.4.2. ToF-SIMS percent was reduced from 35% to 27%.
TOF-SIMS spectra were obtained using a PHI TRIFT II Two additional flotation tests were undertaken at laboratory
System equipped with a gallium liquid metal ion gun (LMIG) in scale, at 30% and 45% solids in the rougher flotation feed.
pulsed mode. For insulating samples the surface charge is These concentra- tions were achieved by manipulating
compensated using a pulsed electron flood gun. The mineral samples of plant feed collected during survey 3 (35% solids),
samples were mounted on to indium foil. The primary beam the former by diluting with process water, the latter by filtering
current employed in the present study was 600 pA (DC and re-suspending the solids in process water. The trend (Fig.
measurement). In static mode, the analysis is confined to the 4d) confirmed what was observed in the previous tests,
top two monolayers. An excitation voltage of 25 kV was used showing high sensitivity of the coarse molybdenite particles to
in un-bunched mode to give a spatial resolution of better than feed solids percent. The feed solids percent is an important
0.5 μm and pulse length adjusted to give a mass resolution (m/ driver to molybdenite recovery overall, and particularly in
Δm) of ~ 4000. Imaging of the sample involved mapping for coarse particle size fractions.
positive and negative ions for surface regions of mineral In Fig. 4, bar charts reporting the distribution of copper and
particles of interest, molybdenum in the feed are superimposed on the recovery
i.e. MoS2. Several tens of particles in each processing stream curves. The area of each bar in the charts is proportional to the
are analysed and the data statistically presented in terms of relative mass of mineral in the specified particle size range.
normalised (to total ion yield) intensity of signals and 95% Considering that, at the standard plant grind, the +150 μm size
confidence intervals. In this mode, statistical differences in fraction contains about 20% of the total molybdenite in the feed,
surface chemistry between the same minerals in different the recovery of this size fraction is critical to the overall
streams or under different pulp conditions may be discerned. In molybdenite recovery.
the present case, we are investigating surface chemistry Surface analysis results on the slurry samples collected
differences between floating and non(slow)-floating MoS 2 during survey 1 (in which the flotation feed consisted mainly
particles. of quartzite type ore, and no limestone skarn was present) are
reported in Fig. 5. No significant difference in surface
3. Results composition between fast (concentrate from Cell 1) and slow
(concentrate from Cell 11) floating molybdenites was found in
3.1. Plant Studies and Laboratory Tests on Plant Slurries this particular case of the quartzite ore type. It could therefore
be concluded that the smaller differences in molybdenite
The grade/recovery relationship for copper and molybdenum recovery noted between plant and laboratory scale (Fig. 3) may
for the four plant surveys and the laboratory flotation tests be ascribed to changes in hydrodynam- ics. However, this
carried out in parallel with the plant surveys are reported in Fig. 3. conclusion will depend on the feed type, as discussed further
The ultimate recovery (combined rougher/scavenger flotation) for below.
copper and molybdenum is also reported in Table 5. In the plant
surveys, molybdenum recovery ranged from 75% to 92%, while 3.2. Laboratory flotation tests on reconstructed feeds
copper recovery from 85% to 92%. The recovery of molybdenum
was consistently lower than copper recovery, the difference The flotation recoveries of copper and molybdenum in
increasing at higher feed solids percent and in the presence of laboratory scale tests on blends of quartzite ore and skarn ore
limestone skarn ore. For both copper and molybdenum, the are reported in Fig. 6. A sharp decrease in molybdenite
lowest recovery was achieved at high solids percent and in the recovery was observed at increasing concentration of
presence of limestone skarn ore (survey 3). Compared to copper, limestone skarn ore in the feed. The ultimate molybdenite
molybdenum recovery was more sensitive to the feed solids recovery decreased from 92% in the absence of skarn ore to
percent, being significantly enhanced when solids percent was 85% with 25% skarn ore and 56% with 50% skarn ore in the
reduced from 35% to 27%. Both copper and molybdenum flotation feed. Copper flotation was affected to a much lesser
recoveries were lower in the presence of limestone extent. This is in agreement with the plant surveys and tests
carried out on 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
particles in the quartzite ore which are depressed in the chemical differences on molybdenite. In the presence of 50%
presence of skarn ore. skarn ore, on the contrary, samples of fast and slow floating
The ToF-SIMS spectra for the fast floating molybdenite (Con molybdenites showed differences in surface chemistry (Fig. 7).
1) and slow floating molybdenite (Con 4) collected in the tests In the fast floating sample (Con 1) there are much more
with 0% and 50% skarn ore in the flotation feed are reported in exposed molybdenum and sulphur compared to the slow
Fig. 7. floating sample (Con 4). Furthermore, Ca, K, Fe, as well as O and
Samples of fast (Con 1) and slow (Con 4) floating OH groups, were more abundant on the latter sample.
molybdenite in the presence of quartzite type ore only, showed Significantly higher Mg on the surface of molybdenite in the
very small differences in surface chemical composition. The presence of skarn ore was also noted in both fast and slow
different species appear in the same proportion on the surface of floating fractions (Fig. 7a).
the two samples, within statistical error (Fig. 7). This is in
agreement with previous 3.3. Laboratory flotation tests on coarse molybdenite particles

Table 5
The effect of calcium and magnesium ions in solution on
Ultimate recovery of copper and molybdenum in plant (after rougher/scavenger coarse molybdenite (+150 μm) recovery from quartzite and
flotation) and laboratory. limestone skarn ores is presented in Fig. 8. In the case of
quartzite ore, the cumulative recovery of molybdenite was
Ore blend Survey Solids Plant Lab recovery
type recovery high, approaching 90% after 8 min of flotation. The addition of
# [%] [%] Cu [%] Mo [%] calcium and magnesium ions to the pulp did not have an effect
Cu [%] Mo
on molybdenite flotation. In contrast, molybdenite recovery
High quartzite 1 27 92 92 95 95
2 35 92 88 94 92
from limestone skarn ore was considerably lower (60%) even in
High skarn 3 35 85 75 90 85 the absence of calcium and magnesium ions, and decreased
4 27 88 85 93 93 further 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 particles. Particle size and shape, morphology and surface
Mg2+ ions and gangue minerals of the skarn ore. composi- tion are all contributing factors. Insufficient
The effect of pH and calcium ions on the contact angle of molybdenite liberation in the coarse size fractions was also
molybdenite was also investigated (Fig. 9). The tests showed considered in the first instance as a possible cause for the low
that faces and edges of molybdenite particles have flotation recovery. However, QEM-Scan analysis of the
considerably different contact angles. Face contact angle was scavenger tailings (Triffett, private communication) showed
high (about 100°) and independent of solution pH and calcium that molybdenite in the +150 μm size fraction mainly consists
concentration. In contrast, contact angle on the edges was low of liberated particles. This confirms that other mechanisms, as
(maximum 45°), and decreased with increased pH and calcium outlined in Section 1.1, play significant roles.
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
4. Discussion molybdenum recovery observed at reduced solids percent (Fig.
3). Higher local turbulent energy dissipation may increase the
From an industrial perspective, the most important finding collision efficiency for the flat and elongated particles, as
from the plant studies was the consistent increase in molybdenite hypothesised in Section 1.2. The fact that molybdenite recovery
recovery in rougher/scavenger flotation at reduced feed solids in laboratory flotation tests was higher than in the plant (Fig. 4)
percent. This is in agreement with previous observations for may also be due hydrodynamic effects. Compared to the plant
different flotation plants (Ametov et al., 2008). The coarse (+ 150 cells, laboratory batch flotation cells have higher local turbulent
μm) size fractions were affected the most by changes in feed energy dissipation (Newell and Grano, 2007), and therefore higher
solids percent, and therefore fluctuations in the overall recovery collision efficiency, Ec, and, according to Eq. (1), higher flotation
of molybdenite in the plant are determined to a large extent by the rate. An interesting feature is that it is the coarse and liberated
flotation response of the coarse 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., scales, suggesting that the different flotation recoveries of
2003). The high surface/bulk ratio of the coarse molybdenite molybde- nite at high and low solids percent may be ascribed to
particles may however be beneficial for the stability of hydrodynamic effects, possibly arising from the intrinsic shape
particle–bubble aggregates, because of a higher surface of the particles, rather than from the surface cleaning of slime
force/inertial force ratio. Surface forces favour attachment of coatings. Therefore, the hypothesis of slime coatings may hold
particles to bubbles, while inertial forces promote detachment. for some feed types and be less important for others. It is
A high surface/bulk ratio results in high stability efficiency, surmised that at low pulp density there is less likelihood of
even in high turbulent conditions such as in laboratory inter-particle interactions causing slime coatings. A reduction
flotation and at reduced solids percent. in slime coatings on the surface of molybdenite particles
Another outcome of the plant studies was that the effect of results in increased contact angle. Even a small increase in
a reduction in feed solids percent on molybdenite recovery was contact angle could cause previously non-floatable coarse
high for some feed blends (high limestone skarn) while it was particles to become floatable and the ultimate recovery is
much lower for others (high quartzite). This was observed at therefore increased.
2−
plant scale (Fig. 3b), in laboratory flotation tests on plant The high concentration of Ca2+ and SO 4 in Kennecott
slurries (Fig. 3d), and in tests on reconstructed feed blends process water (Table 3) may also suggest deposition of calcite
(Fig. 6). Therefore, it is likely that factors other than and gypsum
hydrodynamics and collision efficiency, such as surface from solution due to oversaturation. This could also be a
modification and particle–particle interactions also play a role. contributing factor, but it is not sufficient to justify the different
ToF- SIMS analysis indicated the presence of elements degrees of molybdenite depression observed in the presence of
associated with hydrophilic components on the slow floating quartzite and skarn ores. Diagnostic flotation tests on coarse
coarse molybdenite particles in the presence of skarn ore (Fig. molybdenite particles in the presence of limestone skarn ore and
7). This could be either from the deposition of fine slimes or different levels of calcium and magnesium ions in solution (Fig. 8)
adsorption of dissolved ions. In any case, the effect is strongly corroborate the hypothesis that interactions between
ore type dependent. In the absence of skarn ore, no difference molybdenite and some gangue minerals in the skarn ore lead to
in surface composition between fast and slow molybdenite molybdenite depression. The effect could be induced by metal
was apparent at plant (Fig. 5) and laboratory (Fig. 7) ions adsorbed at the molybdenite edges, as also 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 (pHN
Fig. 8. Recovery of coarse molybdenite particles (+ 150 μm) in flotation tests on 10) and high alkaline conditions ([Ca2+] N 10−2M), as in typical
synthetic feeds. Molybdenite particles from the final concentrate, after surface
flotation environment.
cleaning, were blended with ground quartzite and skarn ore (0.15% MoS2 by
weight). Tests were in demineralised water, at pH 10, in the presence and When limestone skarn ore was blended to the feed in
absence of Ca2+ and Mg2+ ions. laboratory flotation tests, molybdenite recovery decreased
significantly, similarly to what was observed in the plant. Higher
concentrations of Ca, Fe, Mg and K were measured on the slow
faces and edges (Fig. 9). The edges of molybdenite particles floating molybdenite particles compared to the fast floating,
are particularly reactive, and adsorption of Ca2+ and/or Mg2+ which have been correlated to the presence of some gangue
may have the combined effect of reducing the overall minerals typical of the skarn ore. A ‘bridging’ effect of calcium
hydrophobicity of molybdenite particles and “bridging” for and magnesium adsorbed on the surface of molybdenite and
particle–particle interactions with the negatively charged some fine gangue particles in the slurry is possible and may
gangue minerals. This is also in agreement with the findings of lead to the formation of slime coatings, thus reducing the
Raghavan and Hsu (1984) on molybdenite/silica mineral flotation recovery of molybdenite.
systems. Slime coatings are possibly causing molybdenite In terms of strategies to increase molybdenite recovery at
depression in the presence of skarn ore. Minerals which are plant scale, operating the rougher/scavengers at low solid
much more abundant in the limestone skarn ore are: actinolite, percent was a solution which gave immediate benefits. This
andradite, talc and calcite (Table 2). These minerals are benefit may be a result of a combination of hydrodynamic effects
potentially responsible for the high Mg, Ca and Fe observed (higher local turbulent energy dissipation and collision
on the surface of molybdenite when floated in the presence of efficiency), rheology and reduced formation of slime coatings. It
limestone skarn ore (Fig. 7). This conclusion would be also appears that preventing slimes from adsorbing on
consistent with the fact that the presence of the above molybdenite particles during flotation may benefit molybdenite
mentioned minerals in the flotation feed has been correlated to recovery. This is the subject for future investigation.
periods of low molybdenite recovery (Triffett et al., 2008).
AcknowledgeThents

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.
Fig. 9. Contact angle (advancing) of coarse molybdenite particles (+ 150 μm) Hemmings, C.E., 1981. On the significance of flotation froth liquid lamella
measured in proximity of face and edges, after conditioning at different pH thickness.
values and Ca2+ concentrations. Contact angle measured by means of the Trans. Inst. Min. Metall. Sect. C 90, C96–C102.
sessile drop method. Hernlund, R.W., 1961. Extraction of molybdenite from copper flotation products.
Q. J. Colo. Sch. Min. 56, 177–196.
Hoover, M.R., 1980. Water chemistry effects in the flotation of sulphide ores — a
review and discussion for molybdenite. In: Jones, M.J. (Ed.), Complex
Sulphide Ores. IMM, London, pp. 100–112.

You might also like