You are on page 1of 11

Minerals Engineering 83 (2015) 33–43

Contents lists available at ScienceDirect

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

The effect of regrind mills on the separation of chalcopyrite from pyrite


in cleaner flotation
Xumeng Chen, Yongjun Peng ⇑
School of Chemical Engineering, The University of Queensland, St. Lucia, Brisbane, QLD 4072, Australia

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

Article history: Stirred mills have been widely used for regrinding and are more energy efficient than tumbling mills.
Received 6 February 2015 These two types of mills present different particle breakage mechanisms and redox environments during
Revised 11 August 2015 grinding. In this study, the effect of regrinding with these two types of mills on the separation of chal-
Accepted 12 August 2015
copyrite from pyrite in the cleaner stage was studied. A laboratory rod mill and a laboratory stirred mill
Available online 24 August 2015
were used to regrind rougher flotation concentrates. It was found that chalcopyrite and pyrite exhibited
different flotation behavior after regrinding with the rod mill and the stirred mill, resulting in different
Keywords:
separability of chalcopyrite from pyrite. The mechanism underpinning this phenomenon was investi-
Tumbling mill
Stirred mill
gated by a range of techniques including dissolved oxygen demand measurements, X-ray photoelectron
Cleaner flotation spectroscopy (XPS) and Time of flight secondary ion mass spectrometry (ToF-SIMS). It was found that the
Chalcopyrite two mills produced different surface oxidation and pyrite activation by copper ions which determined
Pyrite the separation of chalcopyrite from pyrite. This study demonstrates that the selection of a regrind mill
should not only depend on its energy efficiency but also the property of surfaces produced for subsequent
flotation.
Ó 2015 Elsevier Ltd. All rights reserved.

1. Introduction tumbling mill, impact breakage is predominant, while during the


grinding in a stirred mill, a large proportion of attrition breakage
Regrinding rougher flotation concentrates is typically used to exists especially when particles become fine (Wills and Napier-
liberate valuable minerals from gangue prior to the cleaner stage Munn, 2006; Roufail and Klein, 2010). It has been reported that
in the processing of highly disseminated ores. For example, in breakage mechanisms can influence particle size distribution
Newcrest’s Telfer gold mine, the regrinding of copper rougher con- (Kelly and Spottiswood, 1982; Hogg, 1999), particle shape
centrates significantly improves copper and gold recoveries in the (Vizcarra, 2010) and mineral liberation (Roufail and Klein, 2010;
subsequent flotation (Seaman et al., 2012). The small grain size of Vizcarra et al., 2010) which play an important role in mineral flota-
the valuable minerals in highly disseminated ores requires fine tion. In addition, recent studies have shown that breakage mecha-
grinding to provide sufficient mineral liberation. In some mineral nisms can affect the formation and distribution of mineral surface
processing plants, mineral liberation can only be achieved by species (e.g. collector, contamination species) (Ye et al., 2010a,b;
regrinding to finer than 10 lm (80% passing size) (Johnson, 2006) Vizcarra et al., 2011; Chen et al., 2014a). As presented in Fig. 1, if
while it is becoming increasingly common to regrind to 20 lm. the breakage mechanism is impact, the surface species distribute
Grinding is the single largest energy consuming process in min- on to the surface of particles of all sizes. If attrition breakage is
eral processing plants, and the selection of energy efficient mills is applied, the surface species are removed from the original particle
critical to fine grinding. Stirred mills, recently introduced to min- and tend to distribute on to fine and ultra-fine particles.
eral processing, have been proved to be more energy efficient than Another important aspect of regrinding is the strong electro-
traditional tumbling mills in terms of fine grinding, and have been chemical reactions occurring inside regrind mills. It has been
widely used at the regrinding stage in many mineral processing reported that the interfacial chemical reactions of sulphide miner-
plants (Pease et al., 2006). In addition to energy efficiency, these als during grinding can be affected by different grinding forces
two types of mills also provide different particle breakage mecha- resulting in the formation of crystal lack, dislocation and the nas-
nisms (Kelly and Spottiswood, 1982). During the grinding in a cent surface (Hu et al., 2009). Further, breakage mechanisms not
only directly change the surface chemistry on a single particle,
⇑ Corresponding author. but also change the way of fine particles generated and then indi-
E-mail address: yongjun.peng@uq.edu.au (Y. Peng). rectly change the overall chemical environment in the mill. These

http://dx.doi.org/10.1016/j.mineng.2015.08.008
0892-6875/Ó 2015 Elsevier Ltd. All rights reserved.
34 X. Chen, Y. Peng / Minerals Engineering 83 (2015) 33–43

and then screened to collect +0.71–3.35 mm particle size fractions.


XRD analysis indicated that the pyrite sample was very pure with-
out detecting any impurity. The purity of the chalcopyrite sample
was slightly lower with about 95% chalcopyrite and 5% pyrite.
The processed samples were sealed in polyethylene bags and then
stored in a freezer at a temperature of 20 °C to reduce surface
oxidation.
Potassium amyl xanthate (PAX) and Interfroth 56 were used as
the collector and frother, respectively. They were of industry grade
and used as received. Other chemicals were of AR grade. De-
Fig. 1. Representation of the distribution of surface species during particle breaking ionized water was used in all experiments. Fresh chemical solu-
under impact and attrition mechanisms. tions were prepared daily for flotation tests.

reactions play an important role in the formation of surface spe-


cies. As demonstrated by Chen et al. (2013), there were two types 2.2. Grinding, regrinding and flotation
of surfaces after regrinding pyrite concentrates: 16% of the total
surfaces were the existing surfaces carried from the regrinding The mixture of chalcopyrite (50 g) and pyrite (50 g) was com-
feed, and 84% of the total surfaces were the fresh surfaces gener- bined with 150 g deionized water, and ground in a stainless steel
ated during regrinding. The existing surfaces carried from a rod mill using 4 stainless steel rods (3750 g) to achieve a
regrinding feed were still covered by collector and copper activa- P80 = 75 lm. A 2.5% sodium hydroxide solution was added during
tion species. Both the existing surfaces and freshly produced sur- grinding to achieve pH 9.0 in the primary grinding discharge.
faces can be oxidized during regrinding with an oxidizing After grinding, the pulp was transferred to a flotation cell
environment. Oxidation species formed on surfaces can change (1.5 dm3) for rougher flotation. A JKMRC laboratory batch flotation
the mineral floatability (Smart, 1991; Gonçalves et al., 2003; cell with a bottom-driven agitator was used in this study.
Bicak and Ekmekci, 2012). It is well known that chalcopyrite can Potassium amyl xanthate (160 g/t) and Interfroth 56 (200 g/t) were
be recovered by flotation in the absence of collector under oxidiz- added and 2 min was allowed for the conditioning of each reagent.
ing conditions, but not under reducing conditions (Heyes and During flotation, the pH was maintained at 9.0 by adding a sodium
Trahar, 1977). This is because mild oxidation results in the forma- hydroxide solution (2.5% w/v). The froth was scraped every 10 s,
tion of a hydrophobic sulfur-rich surface due to the dissolution of and four concentrates were collected after cumulative times of
iron ions (Gardner and Woods, 1979; Buckley and Woods, 1984; 0.5, 2.0, 4.0 and 8.0 min. The air flow rate was 1.5 dm3/min during
Zachwieja et al., 1989). In addition, pyrite surface can be activated the first 0.5 min, and then increased to 3.0 dm3/min from the sec-
by copper species (mainly Cu2+ and Cu(OH)2) emanating from cop- ond concentrate.
per bearing minerals/ores during grinding. The interaction The four rougher flotation concentrates were combined and
between Cu2+ and pyrite is an electrochemical process involving mixed with additional water to achieve a pulp density of 18%,
the reduction of Cu2+ to Cu+ with the subsequent oxidation of sur- and then reground in either a rod mill or a stirred mill. The target
face sulphide (Chandra and Gerson, 2009). The produced Cu–S acti- particle size of the regrinding product was P80 = 20 lm. A sodium
vation species can improve the adsorption of xanthate collector by hydroxide solution (2.5%) was added in the feed to achieve pH
forming Cu+–Xanthate. This activation process is highly dependent 9.0 in the regrinding discharge. For regrinding with the rod mill,
on the electrochemical environment. By using voltammetric tech- 10.3 kg stainless steel rods were used, and the particle size
niques, Richardson et al. (1996) found that an increase in oxidizing P80 = 20 lm was achieved after grinding for 15 min. The stirred
potential inhibited copper uptake while an increase in reduction mill used in this study was a vertical bead mill with a disc-type agi-
potential promoted copper uptake. Recent studies also showed tator. It was made by Netszch (model number: M1.5), and the vol-
that a reducing environment during grinding promoted the activa- ume of the grinding chamber was 1.5 L. 1 L ceramic beads were
tion but an oxidizing environment inhibited the activation (Peng used as the grinding media. The diameter of the media was
and Grano, 2010a; Peng et al., 2012; Chen et al., 2013). This phe- 2.5 mm. The regrinding time was 3.5 min at a rotational speed of
nomenon can be explained by the activation mechanism. A reduc- 1200 RPM to achieve the particle size P80 = 20 lm.
ing environment favors the reduction of Cu2+ to Cu+ and hence the The size distribution was determined by Laser Diffraction with
formation of Cu+-sulphide. Overall, the regrind mill can affect min- a Malvern MasterSizer (Malvern Instrument Ltd., U.K.). For the
eral flotation through various electrochemical reactions beyond purpose of measuring size distributions, the regrinding products
particle breakage mechanisms. were filtered and dried at 70 °C in an oven. The dried samples
In this study, the effect of regrind mills on the separation of were split using a micro rotary riffler to produce 0.5 g sub-
chalcopyrite from pyrite in the subsequent cleaner flotation was sample. The sub-sample was mixed with 10 mL water and then
investigated. A laboratory rod mill and a laboratory stirred mill the slurry was placed in an ultrasonic bath for 5 min to disperse
were used to regrind the rougher flotation concentrates. The sur- the fine particles. Then the slurry was transferred to the Laser
face species, including the oxidation species, copper activation spe- Sizer for analysis. Fig. 2 shows the size distribution of the
cies and collector, were analyzed by a number of techniques and regrinding products. More fine particles were produced by the
correlated with the flotation behavior. stirred mill than by the rod mill, which was caused by different
particle breakage mechanisms in these two mills (Kelly and
Spottiswood, 1982).
2. Experimental details After regrinding, the pulp was transferred to a 1.5 dm3 flotation
cell for cleaner flotation. Frother (200 g/t) was added during the
2.1. Materials and reagents conditioning time of 2 min. For some tests, more collector was
added, and the specific amount was detailed in the results. The
Chalcopyrite and pyrite single minerals, supplied by GEO cleaner flotation procedure was the same as used in the rougher
Discoveries, were crushed through a jaw crusher and a roll crusher, flotation after primary grinding.
X. Chen, Y. Peng / Minerals Engineering 83 (2015) 33–43 35

2.3.3. ToF-SIMS study


Time of flight secondary ion mass spectrometry (ToF-SIMS) was
used to study the distribution of surface species on particles of dif-
ferent size fractions after regrinding. The instrument used in this
work was a PHI TRIFT V nanoTOF equipped with a pulsed liquid
metal 79+Au primary ion gun (LMIG), operating at 30 kV energy.
‘‘Unbunched’’ beam settings were used to optimize spatial resolu-
tion. Surface analyses, in positive and negative SIMS modes, were
performed at a number of locations using a 75  75 lm raster area.
For the purpose of statistical interrogation, approximately 25 par-
ticles of interest were imaged per sample to collect representative
data. Region-of-interest analyses were performed on the collected
raw image data, which involved the extraction of mass spectra
specifically from within the boundaries of the particles of interest.
Resulting spectra in each polarity were calibrated using
WincadenceN software (Physical Electronics Inc.) and peaks were
selected based upon previously identified species of interest.
Fig. 2. Particle size distribution (log–log coordinate) after regrinding with the rod
mill and the stirred mill. Integrated peak values of the selected ions were normalized to
the total selected secondary ion intensities, to correct for differ-
ences in total ion yield between analyses and samples. The result-
2.3. Surface analysis ing data were then compared qualitatively by preparing plots of
average normalized counts (with 95% confidence intervals) for
2.3.1. Dissolved oxygen demand measurements the collector species.
The dissolved oxygen (DO) demand measurements were con-
ducted in an airtight 400 mL glass flask, equipped with a magnetic
stirrer to keep the mineral particles in suspension. A YSI dissolved 3. Results and discussion
oxygen probe with a polarographic sensor was used to monitor the
change of dissolved oxygen concentration in the slurry. The probe 3.1. Chemical conditions
was calibrated daily in a sodium sulphite calibration solution and
in air. After turning the device on, it was warmed up for 10 min Chemical conditions inside a grinding mill, including pH, Eh and
before taking readings. The probe was placed in the slurry which dissolved oxygen (DO) concentration, play an important role in the
was stirred all the time and the DO value was recorded automati- electrochemical reactions during grinding and the formation of
cally every minute. 300 mL of the mill discharge obtained after mineral surface species and therefore are critical to the subsequent
regrinding was transferred into the oxygen demand vessel and stir- flotation. In this study, the chemical conditions of the regrinding
red for 1 min to obtain homogeneous pulp after which the initial product were measured immediately after regrinding. The results
DO was recorded. The pulp was then purged with air for 20 min are shown in Table 1. The pH of the regrinding product was
at a rate of 1.5 L/min. After 20 min, the air supply was turned off adjusted to 9 by adding 5.5 mL 2.5% NaOH solution to the rod mill
and oxygen decay was monitored for 30 min. This procedure was regrinding feed and 2.8 mL 2.5% NaOH solution to the stirred mill
repeated several times. The oxygen demand rate constant (k) was regrinding feed. The Eh and DO were higher after regrinding with
then calculated by fitting the first-order rate equation: the rod mill than with the stirred mill. This indicates that the rod
mill produced a more oxidizing grinding condition than the stirred
DO ¼ DO0 ekt mill.
The different chemical conditions should be caused by different
where DO and DO0 are the DO concentrations at time t and t = 0, mill design and different particle breakage mechanisms. In this
respectively. study, the rod mill was operated at a low speed of 76 RPM and
15 min was required to achieve the particle size of P80 = 20 lm.
The stirred mill was more efficient in terms of fine grinding due
2.3.2. XPS analysis
to its high operating speed and small grinding media size. The
The samples for XPS analyses were collected from mill dis-
same P80 was achieved in 3.5 min with the stirred mill operating
charges and also flotation products and were frozen immediately
at 1200 RPM. The longer grinding time indicates a longer reaction
in liquid nitrogen (196 °C). Before characterization, the samples
between minerals and oxygen. However, the shorter grinding of
were thawed and the wet solids were immediately transferred into
the stirred mill still produced a more reducing condition. It is pos-
the high vacuum XPS sample chamber. XPS measurements were
sible that the more efficient generation of fine particles by the
carried out with a KRATOS Axis Ultra (Kratos Analytical,
attrition breakage in a stirred mill consumes oxygen quickly.
Manchester, United Kingdom) with a monochromatic Al X-ray
The chemical conditions in the subsequent cleaner flotation
source operating at 15 kV and 10 mA (150 W). The analysis spot
were also monitored. The pH was controlled to 9 throughout the
area was 300  700 lm. The frozen slurry samples were defrosted
flotation by adding NaOH solution. As shown in Fig. 3, Eh increased
just prior to the analysis. The solids were placed on a stainless
slightly with the flotation and was 10–20 mV lower if the stirred
steel bar and immediately loaded into the introduction chamber
mill was used. The results indicate a similar flotation condition of
of the spectrometer. Samples were analyzed at a pressure of
the two regrinding products.
9  1010 Torr at room temperature. Each analysis started with a
survey scan from 0 to 1200 eV using a pass energy of 160 eV at
Table 1
steps of 1 eV with one sweep. High resolution spectra of Fe 2p, O pH, Eh and DO measured after regrinding with the rod mill and the stirred mill.
1s, C 1s, S 2p, and Cu 2p were collected at 20 eV pass energy at
pH Eh (mV vs SHE) DO (ppm)
steps of 100 meV with two or three sweeps. Binding energies were
charge-corrected by referencing to adventitious carbon at Rod mill 9 324 2.8
Stirred mill 9 278 0.4
284.8 eV.
36 X. Chen, Y. Peng / Minerals Engineering 83 (2015) 33–43

3.2. Flotation results

3.2.1. Rougher flotation


The separation of chalcopyrite from pyrite was studied by using
a mixture of these two minerals. Fig. 4 shows chalcopyrite recovery
as a function of pyrite recovery in rougher flotation after primary
grinding. Both chalcopyrite and pyrite displayed high floatability
with more than 90% recovery at the completion of 8 min flotation.
In the previous study by Chen et al. (2013), under the same flota-
tion condition, pyrite recovery was only 23% without any activa-
tion. However, in the current study, high pyrite recovery was
achieved due to the surface activation of pyrite by copper ions
emanating from chalcopyrite during grinding and conditioning,
improving the collector adsorption on pyrite surfaces at alkaline
conditions (Leppinen, 1990; Finkelstein, 1997). This procedure Fig. 4. Chalcopyrite recovery as a function of pyrite recovery from rougher flotation
was designed to simulate some industrial flotation circuits to max- after primary grinding.
imise copper flotation recovery by floating some pyrite in rougher
flotation and rejecting it in cleaner flotation.

3.2.2. Cleaner flotation


The rougher flotation concentrate was reground in the rod mill
or the stirred mill. The regrinding product was then subjected to
the cleaner flotation stage to study the separation of chalcopyrite
from pyrite. As shown in Fig. 5, after regrinding with the rod mill,
89.0% chalcopyrite recovery was achieved in 8 min flotation with-
out additional collector, while pyrite was strongly depressed with a
recovery of only 9.9%. This indicates that an effective separation of
chalcopyrite from pyrite was achieved in the cleaner stage even
without additional collector. However, the flotation behavior was
significantly different after the same rougher concentrate was
reground with the stirred mill. Chalcopyrite recovery was only
25.7%, while pyrite recovery was 13.3%. Obviously, the separation
of chalcopyrite from pyrite was difficult after regrinding with the
stirred mill.
Chalcopyrite recovery and pyrite recovery from the cleaner Fig. 5. Chalcopyrite recovery as a function of pyrite recovery in the cleaner flotation
without additional collector after regrinding with the rod mill and the stirred mill.
flotation were evaluated on a size-by-size basis and results are
shown in Fig. 6. After regrinding with the rod mill, chalcopyrite
particles presented a high recovery at all size fractions. In compar- entrainment. In addition, recent studies have shown that the
ison, after regrinding with the stirred mill, chalcopyrite recovery chemical properties of particle surfaces also played an important
was low from all size fractions and decreased with an increase in role in flotation. Although fine particles require a lower degree of
particle size. Pyrite recovery was low after regrinding with both hydrophobicity to achieve the same level of floatability as interme-
the rod mill and the stirred mill and was also dependent on particle diate and coarse particles (Trahar, 1981), it has been reported that
size. Finer particles tended to present a higher recovery after the floatability of fine particles were more susceptible to the grind-
regrinding with the rod mill, but the trend was opposite after ing environment (Johnson, 2006; Grano, 2009; Peng and Grano,
regrinding with the stirred mill. From a physical point of view, fine 2010b). The surfaces of fine particles can be more easily contami-
particles have a lower flotation rate due to the low particle-bubble nated by iron species originating from grinding media and can be
collision efficiency and also present a higher degree of more oxidized than intermediate and coarser particles due to the
high surface energy and large surface area. In this study, the signif-
icantly different flotation behavior caused by the two regrind mills
should be more related to the generation of different particle sur-
face properties.
Since a great amount of fresh surfaces was produced after
regrinding due to the size reduction, the collector species carried
from rougher flotation were diluted on the surface. It is reasonable
to expect that additional collector in the cleaner flotation stage will
improve the overall flotation recovery. Flotation results after add-
ing more collector in the cleaner flotation are shown in Fig. 7. After
regrinding with the rod mill, chalcopyrite recovery increased
slightly with the addition of 80 g/t collector but started to decrease
after adding more collector. The decrease in recovery was caused
by the unstable froth which may be related to the high surface
hydrophobicity after more collector adsorbed on the surface
(Johansson and Pugh, 1992). However, pyrite flotation was still
strongly depressed with less than 13% recovery in 8 min flotation
Fig. 3. Eh (vs SHE) measured in the mill discharge and in the flotation cell after
collecting each concentrate.
regardless of the amount of additional collector. The pH was about
X. Chen, Y. Peng / Minerals Engineering 83 (2015) 33–43 37

Fig. 6. Chalcopyrite flotation recovery (a) and pyrite flotation recovery (b) as a function of particle size after regrinding with the rod mill and in the stirred mill.

9 and the Eh was around 300 mV (vs SHE) in the cleaner flotation. 3.3.1. Oxidation species
Under this chemical condition, the xanthate collector was not pos- 3.3.1.1. Oxidation reactions. Oxidation of mineral surfaces occurs
sible to be destroyed. If the collector can adsorb on pyrite surfaces, during regrinding and produces various oxidation species. The oxi-
the floatability of pyrite should be improved. However, pyrite dation of pyrite at alkaline conditions is summarized in reactions 1
recovery was always low regardless of collector addition. This indi- and 2 (Hamilton and Woods, 1981; Zhu et al., 1994). Sulphur or
cates that collector did not adsorb on pyrite surfaces in this system iron-deficient sulphide forms as an intermediate but is rapidly
over the collector addition range tested. further oxidized to sulphate with an increase in pulp potential at
After regrinding with the stirred mill, chalcopyrite recovery was alkaline pH (Hamilton and Woods, 1981; Buckley and Woods,
improved by additional collector in the cleaner flotation, achieving 1987; Zhu et al., 1994).
a maximum recovery of 93% at a collector dosage of 320 g/t. Pyrite
FeS2 þ 3xH2 O ! Fe1x S2 þ xFeðOHÞ3 þ 3xHþ þ 3xe ð1Þ
recovery also increased with collector addition and was 56% at
320 g/t collector. This indicates that the pyrite surfaces produced
by the stirred mill were more reactive with xanthate collector Fe1x S2 þ ð11  3xÞH2 O ! ð1  xÞFeðOHÞ3 þ 2SO2
4

compared to the surfaces produced by the rod mill. Since both þ ð19  3xÞHþ þ ð15  3xÞe ð2Þ
chalcopyrite and pyrite recoveries increased in the presence of
where 0 < x 6 1.
additional collector, the separation of chalcopyrite from pyrite
The oxidation of chalcopyrite at alkaline conditions is shown in
was still difficult after regrinding with the stirred mill.
reaction 3. Similar to pyrite oxidation, chalcopyrite oxidation also
In conclusion, the separation of chalcopyrite from pyrite was
starts from the release of iron ions which form iron oxidation spe-
different after regrinding with the rod mill and the stirred mill.
cies on the surface. As a result, a sulphur-rich surface is produced,
High chalcopyrite recovery was observed after regrinding with
although there remains an argument whether it is a metal-
the rod mill, while pyrite flotation was strongly depressed and
deficient sulphide lattice, a metal polysulphide or elemental
could not be restored by additional collector. Therefore, a high sep-
sulphur (Gardner and Woods, 1979; Buckley and Woods, 1984).
aration of chalcopyrite from pyrite was achieved. By contrast, after
Further oxidation of CuFe1xS2 also occurs after extended exposure
regrinding with the stirred mill, both chalcopyrite and pyrite flota-
to air producing copper oxidation species and sulphate (Vaughan
tion were low without additional collector in the cleaner flotation,
et al., 1997).
but the recovery of both minerals increased simultaneously after
adding more collector. Therefore, the separation of chalcopyrite CuFeS2 þ 3xH2 O ! CuFe1x S2 þ xFeðOHÞ3 þ 3xHþ þ 3xe ð3Þ
from pyrite was difficult after regrinding with the stirred mill.
In this study, the particle size of the two regrinding products where 0 < x 6 1.
and the flotation conditions were similar, therefore, the signifi-
cantly different flotation behavior after regrinding with different 3.3.1.2. Dissolved oxygen (DO) demand measurements. The dissolved
mills should be more related to the generation of different particle oxygen (DO) demand measurements were applied to understand
surface properties, i.e. surface oxidation, copper activation on pyr- the overall surface oxidation of regrinding products (Spira and
ite and the distribution of collector. This was investigated by Rosenblum, 1978). Results are shown in Fig. 8. The air purging into
applying different surface analysis techniques and was detailed slurry was turned on for 20 min and then turned off for 30 min.
in the following section. This was repeated 3 times. The DO level increased to a maximum
after purging air for some time. After turning off the air, DO started
to decrease due to the consumption of oxygen in the slurry by min-
3.3. Surface species eral oxidation. The oxygen consumption rate of each cycle was cal-
culated and summarized in Table 2. After air purging for 20 min,
Two types of surfaces were produced after regrinding. One is the rate constant k was 0.065 for the rod mill regrinding product
the remaining surfaces carried over from the regrinding feed which and 0.225 for the stirred mill regrinding product. The rate constant
were covered by the species formed during primary grinding and decreased after more aeration. A higher rate constant indicates a
rougher flotation, such as collector, oxidation species and copper more reactive surface. The DO demand measurements in this study
activation species. As studied by Chen et al. (2014a), the distribu- clearly show that the mineral surfaces produced by the rod mill
tion of these species on regrinding products was influenced by dif- were more oxidized compared to the surfaces produced by the stir-
ferent particle breakage mechanisms provided by different regrind red mill.
mills, which resulted in different mineral floatability in the cleaner
stage. The other type of surfaces are the fresh ones generated dur- 3.3.1.3. XPS analysis. XPS was applied to quantitatively analyse the
ing regrinding. These fresh surfaces can be easily oxidized or con- surface species of the regrinding products. The XPS-measured
taminated during regrinding due to the fine size, large surface area atomic concentrations normalized to oxygen, sulphur, iron, and
and strong galvanic interactions (Chen et al., 2014b,c). The surface copper on the mineral surfaces are given in Table 3 and represent
species of both types were examined in this study. the percentage of surface coverage by different elements. 35.4% of
38 X. Chen, Y. Peng / Minerals Engineering 83 (2015) 33–43

Fig. 7. Chalcopyrite (a) and pyrite (b) recovery from cleaner flotation as a function of collector dosage.

Table 3
XPS derived surface concentrations (atomic % normalized to Fe, S, O and Cu) on the
surfaces of flotation rougher concentrates before and after regrinding with different
grinding mills.

Samples Atomic concentration of element (at.%)


O Fe S Cu
Rod mill 35.4 16.6 36.2 11.7
Stirred mill 27.7 16.0 43.9 12.4

especially the peak of S22 from pyrite and the peak of S22 from
chalcopyrite, it is difficult to further decouple them precisely
(Smart et al., 1999). An energy loss (EL) feature was fitted at
approximately 164.4 eV. It is associated with S 3p to Fe 3d excita-
tion (Harmer et al., 2004; Acres et al., 2010). Furthermore, a broad
peak was observed at 168.2 eV on the rod mill regrinding product,
representing the sulphate SO2 4 (Buckley and Woods, 1984). The
quantification of sulphur of both regrinding feed and product
Fig. 8. DO level as a function of time for the regrinding product from the rod mill
was summarized in Table 4. After regrinding with the rod mill,
and the stirred mill.
26.4 at% of S occurred in the form of S2, 61.5 at% of S was present

the surface elements were oxygen after regrinding with the rod
mill, while this percentage decreased to 27.7% after regrinding
with the stirred mill. The higher oxygen concentration after
regrinding with the rod mill indicates that the mineral surfaces
were covered by more oxidation species. In addition, more sulphur
species were detected on the stirred mill product, corresponding to
the lower surface oxidation. The result is consistent with the dis-
solved oxygen demand profile.
It is well known that collectorless flotation of chalcopyrite
occurs under mildly oxidizing conditions due to the formation of
a hydrophobic sulphur-rich surface (Heyes and Trahar, 1977). By
contrast, the sulphate species are hydrophilic, and can depress
the mineral flotation (Smart, 1991). In this study, the sulphur oxi-
dation species on mineral surfaces were investigated by S 2p XPS
spectra. The results are shown in Fig. 9. The spectra were fitted
with two doublets. The peak at 161.3 eV is attributable to S2 from
chalcopyrite or pyrite. The other doublet at 162.5 eV is due to the
combination of S22 from pyrite, S2
2 from chalcopyrite, and Sn2
from chalcopyrite. Since the positions of these peaks are very close,

Table 2
The first order rate constant of oxygen consumption after different aeration times.

The first order rate constant k (min1)


20 min aeration 40 min aeration 60 min aeration
Rod mill 0.065 0.053 0.046
Stirred mill 0.225 0.142 0.059 Fig. 9. S 2p XPS spectra of mineral surfaces of regrinding products from the rod mill
(top) and the stirred mill (bottom).
X. Chen, Y. Peng / Minerals Engineering 83 (2015) 33–43 39

as S2 2
2 and 4.8 at% S was detected in the form of SO4 . The presence with the two regrind mills should be related to the specific particle
2
of SO4 species on the surfaces indicates strong oxidation occur- breakage mechanism or redox environment. Regrinding with the
ring during regrinding. In comparison, a higher percentage of S pre- stirred mill generated more fine particles than the rod mill with
sented as S2 and S2 2 on the stirred mill regrinding product and no the same P80 as presented in Fig. 2, which indicates a higher speci-
SO2
4 species were detected. It suggests that less oxidation of sul- fic surface area of the stirred mill regrinding product. In terms of
phur occurred on mineral surfaces after regrinding with the stirred redox environment, the rod mill produced a more oxidizing condi-
mill. tion than the stirred mill as shown in Table 1. Therefore, the degree
In this study, as chalcopyrite and pyrite were mixed in the of mineral oxidation in the rod mill was higher than in the stirred
regrinding product, it was difficult to distinguish chalcopyrite oxi- mill. In addition, there was no statistically significant difference
dation species from pyrite oxidation species through XPS analysis. among the particles at different size fractions in all cases in Fig. 11.
When the two minerals were well separated in the cleaner stage, The sulphur species on chalcopyrite and pyrite were also inves-
the examination of flotation concentrates and tailings is an alterna- tigated by ToF-SIMS. Fig. 12 shows the intensity of the ToF-SIMS
tive approach to identifying the surface species on each mineral. As secondary ions of total sulphur (S) on each mineral. After regrind-
discussed previously, high chalcopyrite recovery and low pyrite ing with the rod mill, the intensity of the S signal on chalcopyrite
recovery was obtained in the cleaner flotation after regrinding with surfaces was significantly higher than that on pyrite surfaces.
the rod mill. From the chemical assay, the first cleaner flotation Based on the XPS analysis, the high S intensity on chalcopyrite sur-
concentrate consisted of 99.8% chalcopyrite, while the flotation face may be caused by the oxidation of the surface which formed
tailing consisted of 96.1% pyrite. Therefore, the XPS analyses on metal-deficient sulphide or polysulphide due to the dissolution
the first flotation concentrate and tailing were conducted to under- of Fe or Cu ions. These hydrophobic species contributed to the high
stand the oxidation species on chalcopyrite and pyrite, respec- chalcopyrite floatability after regrinding. The sulphur intensity on
tively. Fig. 10 shows the S 2p XPS spectra of the first concentrate pyrite surfaces was lower. This is consistent with the XPS results
and tailing. The quantification of these sulphur species is summa- which showed that the sulphate oxidation species mainly formed
rized in Table 5. On chalcopyrite surfaces in the flotation concen- on pyrite rather on chalcopyrite surfaces. However, after regrind-
trate, 19 at% sulphur occurred as hydrophobic polysulphide and ing with the stirred mill, the difference in sulphur intensity on
no hydrophilic sulphate species were detected. It indicates that chalcopyrite and pyrite became smaller. This indicates the
highly hydrophobic surfaces were produced on chalcopyrite after decreased oxidation on both mineral surfaces. Less hydrophobic
regrinding with the rod mill, contributing to the high chalcopyrite sulphide species formed on chalcopyrite, resulting in a low recov-
flotation recovery in the cleaner stage. On pyrite surfaces in the ery after regrinding. Meanwhile, less sulphate species formed on
flotation tailing, a clear SO2
4 peak at 168.2 eV was detected with pyrite surfaces, which can benefit pyrite flotation in the cleaner
a percentage of 11.1 at% of total S, suggesting strong oxidation of flotation.
pyrite surfaces during regrinding. These hydrophilic sulphate spe- Overall, the surface analysis show that mineral surfaces were
cies can depress mineral flotation and play a role in the low pyrite more oxidized after regrinding with the rod mill than with the stir-
recovery in the cleaner flotation. red mill. The high chalcopyrite recovery after regrinding with the
rod mill was due to the sulphur rich oxidation species formed on
the surfaces. A significant amount of SO2 4 species was detected
3.3.1.4. ToF-SIMS. To further understand the surface oxidation spe- on the rod mill regrinding product and mainly distributed on pyrite
cies on chalcopyrite and pyrite surfaces and also their distribution surfaces, reducing the surface hydrophobicity. Therefore, a high
on particles of different size factions, ToF-SIMS analysis was per- separation of chalcopyrite from pyrite was achieved. However,
formed on regrinding products. The information for the different low surface oxidation occurred during regrinding with the stirred
mineral phases in the samples was extracted individually for chal- mill and the chalcopyrite was not sufficiently oxidised to produce
copyrite and pyrite particles at the size fraction of 10 lm, 10– hydrophobic sulphur rich species. This could be the main reason
20 lm and +20 lm. for the low chalcopyrite recovery. Meanwhile, pyrite oxidation
Fig. 11 shows the intensity of the ToF-SIMS secondary ions of was also low with more fresh surfaces created. Pyrite was recov-
total oxygen (O) on chalcopyrite and pyrite particles at different ered in the cleaner flotation with the addition of more collector
size fractions after regrinding with the rod mill and stirred mill. resulting in a difficult separation of chalcopyrite from pyrite.
On the rod mill regrinding product, the intensity of oxygen signals
was higher on pyrite surfaces than on chalcopyrite surfaces. By 3.3.2. Copper activation species on pyrite surfaces
contrast, after regrinding with the stirred mill, the intensity of oxy- In this study, pyrite was activated during primary grinding by
gen signals on chalcopyrite and pyrite surfaces was similar. This copper ions emanating from chalcopyrite, leading to high pyrite
indicates stronger pyrite oxidation during regrinding with the recovery in the subsequent rougher flotation. During regrinding,
rod mill than with the stirred mill, which is consistent with the in addition to the activation species carried over from rougher con-
XPS analysis. The different mineral oxidation during regrinding centrates, the generated fresh surfaces were possible to be further
activated under the suitable chemical environment inside the
grinding mill. The activation species on pyrite surfaces can increase
Table 4
S 2p quantification of regrinding products.
the adsorption of collector on pyrite and improve the subsequent
pyrite cleaner flotation. The flotation results in Fig. 7 show that
Sample Species 2p3/2 Area contribution
pyrite flotation recovery was restored by additional collector after
position (%)
regrinding with the stirred mill, however, after regrinding with the
Rod mill Monosulphide 161.3 26.4 rod mill, pyrite recovery was difficult to be restored even in the
Disulphide/ 162.5 61.5
Polysulphide
presence of a high dosage of collector. It is possible that the differ-
Sulphate 168.2 4.8 ent flotation behavior of pyrite after regrinding with the two mills
Energy loss 164.4 7.3 was caused by the different coverage of copper activation species.
Stirred Monosulphide 161.2 27.3 Surface analysis was applied to investigate the formation and
mill Disulphide/ 162.5 67.7 distribution of copper activation species. Although there are still
Polysulphide some debates about the exact activation species, it is generally
Energy loss 164.3 5.1
agreed that the activation on pyrite surface is mainly due to the
40 X. Chen, Y. Peng / Minerals Engineering 83 (2015) 33–43

Fig. 10. S 2p XPS spectra of mineral surfaces of the first cleaner flotation concentrate (a) and flotation tailing (b) after regrinding with the rod mill.

Table 5 Cook, 1975; Deroubaix and Marcus, 1992). A second component


S 2p quantification of the first cleaner flotation concentrate and flotation tailing after in the Cu 2p3/2 at about 934.0 eV is attributable to Cu2+-hydroxyl
regrinding with the rod mill.
species (Weisener and Gerson, 2000). The percentage of Cu+ to
Sample Species 2p3/2 Area contribution the overall Cu for the rod mill regrinding product and the stirred
position (%) mill regrinding product was 84% and 85%, respectively. In both
Flotation Monosulphide 161.3 57.8 regrinding products, Cu2+ ions were the minor copper species on
concentrate Disulphide 162.3 16.9 mineral surfaces. However, the XPS analysis measures the total
Polysulphide 162.7 19.0
copper species on both chalcopyrite and pyrite, therefore the dis-
Energy loss 164.3 6.4
tribution of copper species on pyrite cannot be determined only
Flotation tailing Monosulphide 161.5 10.5
by XPS.
Disulphide 162.5 65.6
Polysulphide 163.4 6.6 To further examine the distribution of copper species on pyrite
Sulphate 168.3 11.1 in different regrinding products, ToF-SIMS was applied to measure
Energy loss 164.3 6.2 the intensity of copper ions on both minerals. As shown in Fig. 14,
after regrinding with both mills, a certain amount of copper ions
was detected on pyrite surfaces, which indicates the existence of
reduction of Cu2+ to Cu+ and the formation of a new copper sul- copper activation species. As discussed in XPS results, more than
phide phase while S2 1
2 is oxidized to S2 or Sn
2
. Therefore, in this 80% copper ions presented in the form of Cu+. Therefore, a signifi-
study, the valence state of copper species was firstly studied by cant percentage of copper species on pyrite surfaces should be Cu+
XPS. The Cu 2p XPS spectra from mineral surfaces prepared by species which were either carried from rougher flotation or newly
the rod mill and stirred mill are shown in Fig. 13. The Cu+ Cu produced during regrinding. Although the intensity of copper ions
2p3/2 component was identified at about 932.2 eV (McIntyre and on chalcopyrite surfaces was similar after regrinding with both

Fig. 11. ToF-SIMS normalised intensity of oxygen on particle surfaces of both chalcopyrite and pyrite at different size fractions after regrinding with the rod mill (a) and the
stirred mill (b).
X. Chen, Y. Peng / Minerals Engineering 83 (2015) 33–43 41

Fig. 12. ToF-SIMS normalised intensity of sulphur on chalcopyrite and pyrite at different size fractions after regrinding with the rod mill (a) and stirred mill (b).

stirred mill. In addition, no statistical difference of the intensity


of copper ions was identified on the particles at different size frac-
tions. Among all these copper activation species, some were carried
over from the rougher concentrates, while the rest was produced
during regrinding. However, it is difficult to distinguish the exact
source of these species. As discussed previously, in the rod mill,
the main breakage mechanism is impact, and the carried species
can remain on the surface of particles of all sizes. In the stirred mill,
the attrition breakage predominates and tends to remove the car-
ried species from surfaces on to fine and ultra-fine particles.
Therefore, in this study, the excess copper species on the stirred
mill regrinding product may be mainly produced during regrind-
ing. In addition, these two mills produced different surface oxida-
tion. More oxidation occurred after regrinding with the rod mill
than with the stirred mill, especially on pyrite surfaces. As studied
by Chen et al. (2014b), the pyrite activation by copper ions can be
prohibited by surface oxidation, therefore, copper activation rate of
Fig. 13. Cu 2p XPS spectra from chalcopyrite-pyrite mixture after regrinding with
the stirred mill (top) and the rod mill (bottom).
pyrite during regrinding with the rod mill was lower than that dur-
ing regrinding with the stirred mill. The insufficient copper activa-
tion after regrinding with the rod mill resulted in low pyrite
mills, a significantly higher intensity was found on pyrite surfaces recovery in the cleaner flotation even after adding more collector.
after regrinding with the stirred mill. This suggests that more cop- By contrast, less oxidation occurred on pyrite surfaces during
per activation species existed on pyrite surfaces produced by the regrinding with the stirred mill, promoting the pyrite activation

Fig. 14. ToF-SIMS normalised intensity of copper ions on particle surfaces of both chalcopyrite and pyrite at different size fractions after regrinding with the rod mill (a) and
stirred mill (b).
42 X. Chen, Y. Peng / Minerals Engineering 83 (2015) 33–43

Fig. 15. ToF-SIMS normalised intensity of collector fragment (potassium amyl xanthate) on chalcopyrite and pyrite at different size fractions after regrinding with the rod
mill (a) and the stirred mill (b).

by copper ions and improving pyrite flotation recovery in the pres- The surface analysis showed that the different flotation behav-
ence of a sufficient amount of collector. ior after regrinding with the rod mill and the stirred mill was
mainly caused by different oxidation species formed during
3.3.3. Collector species regrinding and different copper activation on pyrite surfaces.
Collector species were carried over from the rougher flotation Mineral surfaces were more oxidized after regrinding with the
and distributed on the surfaces of newly produced particles after rod mill than with the stirred mill. Oxidation of chalcopyrite pro-
regrinding. The distribution of these collector species was analyzed duced a hydrophobic sulphur-rich surface which promoted collec-
by ToF-SIMS. As shown in Fig. 15, after regrinding with the rod torless flotation of chalcopyrite and led to high chalcopyrite
mill, more collector species were detected on pyrite surfaces than recovery in the cleaner stage. Oxidation of pyrite produced a signif-
on chalcopyrite surfaces. The collector concentration on the parti- icant amount of hydrophilic species (e.g. iron hydroxide, sulphate)
cles at different size fractions was similar. Although an overall on the surface, which not only reduced its surface hydrophobicity
higher concentration of collector remained on pyrite surfaces than but also prohibited its activation by copper ions, resulting in low
on chalcopyrite surfaces, the collector species did not provide suf- pyrite recovery even in the presence of additional collector.
ficient surface hydrophobicity for pyrite flotation partly because of Therefore, a high separation of chalcopyrite from pyrite was
the increased total surface area diluting the collector concentra- achieved in the cleaner stage after regrinding with the rod mill.
tion. High chalcopyrite recovery was achieved after regrinding By contrast, low surface oxidation occurred during regrinding with
with the rod mill, which was more likely to be caused by the the stirred mill. Chalcopyrite was not sufficiently oxidised to pro-
hydrophobic sulphur species rather than the collector species car- duce hydrophobic sulphur rich species, which was the main reason
ried over from regrinding feed. After regrinding with the stirred for the low chalcopyrite recovery. Meanwhile, low pyrite surface
mill, the collector distribution on particles at different size frac- oxidation benefitted the copper activation of pyrite. The activated
tions was different. A higher concentration of collector was pyrite could be recovered in the cleaner flotation with the addition
detected on chalcopyrite particles larger than 20 lm than on finer of more collector and also competed for collector with chalcopy-
chalcopyrite particles. The trend was opposite for pyrite and sur- rite, resulting in a difficult separation of these two minerals.
face collector concentration decreased as particle size increased.
Different collector distribution on minerals may be related to the
different breakage mechanisms. In the subsequent cleaner flota- Acknowledgments
tion, both chalcopyrite and pyrite flotation were strongly
depressed with a low recovery. Therefore, the collector carried The authors gratefully acknowledge financial support from
from rougher flotation was diluted due to the increased surface Newcrest Mining Limited and the Australian Research Council,
area and their influences on mineral flotation in the cleaner stage and the facilities, and scientific and technical assistance of the
became less significant. Centre for Microscopy and Microanalysis (CMM) at the
University of Queensland and the Australian Microscopy &
Microanalysis Research Facility at the South Australian Regional
4. Conclusions
Facility (SARF), University of South Australia. The first author also
thanks the scholarship provided by the University of Queensland.
The current study indicates that different regrind mills have a
significant influence on the separation of chalcopyrite from pyrite
in the cleaner flotation. High chalcopyrite recovery was observed References
after regrinding with a rod mill, while pyrite flotation was strongly
depressed and could not be restored by additional collector, lead- Acres, R.G., Harmer, S.L., Beattie, D.A., 2010. Synchrotron XPS studies of solution
exposed chalcopyrite, bornite, and heterogeneous chalcopyrite with bornite.
ing to a high separation of chalcopyrite from pyrite. By contrast,
Int. J. Miner. Process. 94 (1–2), 43–51.
after regrinding with a stirred mill, both chalcopyrite recovery Bicak, O., Ekmekci, Z., 2012. Prediction of flotation behavior of sulphide ores by
and pyrite recovery were low without additional collector in the oxidation index. Miner. Eng. 36–38, 279–283.
cleaner flotation, and the recovery of both minerals was improved Buckley, A., Woods, R., 1984. An X-ray photoelectron spectroscopic study of the
oxidation of chalcopyrite. Aust. J. Chem. 37 (12), 2403–2413.
simultaneously after adding more collector, resulting in a low sep- Buckley, A.N., Woods, R., 1987. The surface oxidation of pyrite. Appl. Surf. Sci. 27 (4),
aration of chalcopyrite from pyrite. 437–452.
X. Chen, Y. Peng / Minerals Engineering 83 (2015) 33–43 43

Chandra, A.P., Gerson, A.R., 2009. A review of the fundamental studies of the copper Peng, Y., Grano, S., 2010b. Inferring the distribution of iron oxidation species on
activation mechanisms for selective flotation of the sulfide minerals, sphalerite mineral surfaces during grinding of base metal sulphides. Electrochim. Acta 55
and pyrite. Adv. Colloid Interface Sci. 145 (1–2), 97–110. (19), 5470–5477.
Chen, X., Peng, Y., Bradshaw, D., 2013. Effect of regrinding conditions on pyrite Peng, Y., Wang, B., Gerson, A., 2012. The effect of electrochemical potential on the
flotation in the presence of copper ions. Int. J. Miner. Process. 125, 129–136. activation of pyrite by copper and lead ions during grinding. Int. J. Miner.
Chen, X., Peng, Y., Bradshaw, D., 2014a. The effect of particle breakage mechanisms Process. 102–103, 141–149.
during regrinding on the subsequent cleaner flotation. Miner. Eng. 66–68, 157– Richardson, P.E., Chen, Z., Tao, D.P., Yoon, R.H., 1996. Electrochemical control of
164. pyrite activation by copper. In: Woods, R., Doyle, F., Richardson, P.E. (Eds.),
Chen, X., Peng, Y., Bradshaw, D., 2014b. The separation of chalcopyrite and Electrochemistry in Mineral and metal Processing. Electrochem. Soc.,
chalcocite from pyrite in cleaner flotation after regrinding. Miner. Eng. 58, 64– Pennington, USA, pp. 179–190.
72. Roufail, R., Klein, B., 2010. Mineral liberation and particle breakage in stirred mills.
Chen, X., Seaman, D., Peng, Y., Bradshaw, D., 2014c. Importance of oxidation during Can. Metall. Q. 49 (4), 419–428.
regrinding of rougher flotation concentrates with a high content of sulphides. Seaman, D.R., Burns, F., Adamson, B., Manton, P., 2012. Telfer processing plant
Miner. Eng. 66–68, 165–172. upgrade – the implementation of additional cleaning capacity and the
Deroubaix, G., Marcus, P., 1992. X-ray photoelectron spectroscopy analysis of regrinding of copper and pyrite concentrates. In: 11th AusIMM Mill
copper and zinc oxides and sulphides. Surf. Interface Anal. 18 (1), 39–46. Operators’ Conference, Hobart, Tasmania.
Finkelstein, N.P., 1997. The activation of sulphide minerals for flotation: a review. Smart, R.S.C., 1991. Surface layers in base metal sulphide flotation. Miner. Eng. 4 (7–
Int. J. Miner. Process. 52 (2–3), 81–120. 11), 891–909.
Gardner, J.R., Woods, R., 1979. An electrochemical investigation of the natural Smart, R.S.C., Skinner, W.M., Gerson, A.R., 1999. XPS of sulphide mineral surfaces:
floatability of chalcopyrite. Int. J. Miner. Process. 6 (1), 1–16. metal-deficient, polysulphides, defects and elemental sulphur. Surf. Interface
Gonçalves, K.L.C., Andrade, V.L.L., Peres, A.E.C., 2003. The effect of grinding Anal. 28 (1), 101–105.
conditions on the flotation of a sulphide copper ore. Miner. Eng. 16 (11), Spira, P., Rosenblum, F., 1978. Application of oxygen demand to aeration. Can. Min.
1213–1216. J. (March), 28–29
Grano, S., 2009. The critical importance of the grinding environment on fine particle Trahar, W.J., 1981. A rational interpretation of the role of particle size in flotation.
recovery in flotation. Miner. Eng. 22 (4), 386–394. Int. J. Miner. Process. 8 (4), 289–327.
Hamilton, I.C., Woods, R., 1981. An investigation of surface oxidation of pyrite and Vaughan, D.J., Becker, U., Wright, K., 1997. Sulphide mineral surfaces: theory and
pyrrhotite by linear potential sweep voltammetry. J. Electroanal. Chem. 118 experiment. Int. J. Miner. Process. 51 (1–4), 1–14.
(Journal Article), 327–343. Vizcarra, T., 2010. The Effect of Comminution Mechanism on Particle Properties:
Harmer, S.L., Pratt, A.R., Nesbitt, W.H., Fleet, M.E., 2004. Sulfur species at Consequences for Downstream Flotation Performance. PhD Thesis, The
chalcopyrite (CuFeS2) fracture surfaces. Am. Mineral. 89 (7), 1026–1032. University of Queensland.
Heyes, G.W., Trahar, W.J., 1977. The natural flotability of chalcopyrite. Int. J. Miner. Vizcarra, T.G., Harmer, S.L., Wightman, E.M., Johnson, N.W., Manlapig, E.V., 2011.
Process. 4 (4), 317–344. The influence of particle shape properties and associated surface chemistry on
Hogg, R., 1999. Breakage mechanisms and mill performance in ultrafine grinding. the flotation kinetics of chalcopyrite. Miner. Eng. 24 (8), 807–816.
Powder Technol. 105 (1), 135–140. Vizcarra, T.G., Wightman, E.M., Johnson, N.W., Manlapig, E.V., 2010. The effect of
Hu, Y., Sun, W., Wang, D., 2009. Electrochemistry of flotation of sulphide minerals. breakage mechanism on the mineral liberation properties of sulphide ores.
Springer, New York. Miner. Eng. 23 (5), 374–382.
Johansson, G., Pugh, R.J., 1992. The influence of particle size and hydrophobicity on Weisener, C., Gerson, A., 2000. Cu(II) adsorption mechanism on pyrite: an XAFS and
the stability of mineralized froths. Int. J. Miner. Process. 34 (1–2), 1–21. XPS study. Surf. Interface Anal. 30 (1), 454–458.
Johnson, N.W., 2006. Liberated 0–10 lm particles from sulphide ores, their Wills, B.A., Napier-Munn, T.J., 2006. Wills’ Mineral Processing Technology: An
production and separation – recent developments and future needs. Miner. Introduction to the Practical Aspects of Ore Treatment and Mineral Recovery.
Eng. 19 (6–8), 666–674. Butterworth-Heinemann, Oxford.
Kelly, Z.G., Spottiswood, D.J., 1982. Introduction to Mineral Processing. Wiley, New Ye, X., Gredelj, S., Skinner, W., Grano, S.R., 2010a. Regrinding sulphide minerals –
York. breakage mechanisms in milling and their influence on surface properties and
Leppinen, J.O., 1990. FTIR and flotation investigation of the adsorption of ethyl flotation behaviour. Powder Technol. 203 (2), 133–147.
xanthate on activated and non-activated sulfide minerals. Int. J. Miner. Process. Ye, X., Gredelj, S., Skinner, W., Grano, S.R., 2010b. Evidence for surface cleaning of
30 (3–4), 245–263. sulphide minerals by attritioning in stirred mills. Miner. Eng. 23 (11–13), 937–
McIntyre, N.S., Cook, M.G., 1975. X-ray photoelectron studies on some oxides and 944.
hydroxides of cobalt, nickel, and copper. Anal. Chem. 47 (13), 2208–2213. Zachwieja, J.B., McCarron, J.J., Walker, G.W., Buckley, A.N., 1989. Correlation
Pease, J.D., Curry, D.C., Young, M.F., 2006. Designing flotation circuits for high fines between the surface composition and collectorless flotation of chalcopyrite. J.
recovery. Miner. Eng. 19 (6–8), 831–840. Colloid Interface Sci. 132 (2), 462–468.
Peng, Y., Grano, S., 2010a. Effect of grinding media on the activation of pyrite Zhu, X., Li, J., Wadsworth, M.E., 1994. Characterization of surface layers formed
flotation. Miner. Eng. 23 (8), 600–605. during pyrite oxidation. Colloids Surf., A 93, 201–210.

You might also like