You are on page 1of 14

Int. J. Miner. Process.

69 (2003) 87 – 100
www.elsevier.com/locate/ijminpro

Control of grinding conditions in the flotation of


chalcopyrite and its separation from pyrite
Yongjun Peng a, Stephen Grano b,*, Daniel Fornasiero b, John Ralston b
a
COREM, 1180, rue de la Minéralogie, Quebec City, QC, Canada G1N 1X7
b
Ian Wark Research Institute, The ARC Special Research Centre for Particle and Material Interfaces,
University of South Australia, Mawson Lakes, Adelaide, SA 5095, Australia

Received 23 April 2002; accepted 10 September 2002

Abstract

A specially designed mill which allowed the control of pH throughout grinding was used to study
the effect of grinding conditions on chalcopyrite flotation and chalcopyrite separation from pyrite.
The mechanism of galvanic interaction between minerals and grinding media was investigated by
ethylene diamine-tetra acetic acid disodium salt (EDTA) extraction and X-ray photoelectron
spectroscopy (XPS) measurements.
Chalcopyrite flotation was strongly dependent on both iron oxidation species and metal
deficiency on the chalcopyrite surface. Iron oxidation species from grinding media played a
dominant role in depressing chalcopyrite flotation, while metal deficiency from chalcopyrite
oxidation improved chalcopyrite flotation. Therefore, chromium grinding medium produced a higher
chalcopyrite recovery than mild steel grinding medium while gas purging during grinding had little
effect on chalcopyrite flotation.
Chalcopyrite separation from pyrite was affected by the activation of pyrite flotation by copper
species dissolved from chalcopyrite. Grinding media had a large effect on the reduction of copper(II)
to copper(I) on the pyrite surface. The reducing grinding condition generated by mild steel medium
favoured formation of copper(I) sulphide phase, which resulted in high pyrite activation. Thus,
chromium medium produced better chalcopyrite selectivity against pyrite than the mild steel
medium.
D 2003 Elsevier Science B.V. All rights reserved.

Keywords: chalcopyrite; pyrite; grinding; flotation; copper activation; galvanic coupling

* Corresponding author. Fax: +61-8-8302-3683.


E-mail address: stephen.grano@unisa.edu.au (S. Grano).

0301-7516/03/$ - see front matter D 2003 Elsevier Science B.V. All rights reserved.
PII: S 0 3 0 1 - 7 5 1 6 ( 0 2 ) 0 0 11 9 - 9
88 Y. Peng et al. / Int. J. Miner. Process. 69 (2003) 87–100

1. Introduction

It has been known for a long time that grinding conditions have a significant influence
on the subsequent flotation of sulphide minerals. For chalcopyrite flotation, there is
agreement that mild steel grinding medium results in a lower chalcopyrite recovery than
stainless steel grinding medium in the presence or absence of collectors (Rao et al., 1976;
Forssberg et al., 1988; Van Deventer et al., 1991). The depression of chalcopyrite flotation
by mild steel medium was attributed to the iron hydroxide species on chalcopyrite surface,
which was closely related to the anodic iron oxidation (Yelloji Rao and Natarajan, 1988).
The rate of anodic iron oxidation was found to decrease by nitrogen purging during wet
grinding of chalcopyrite minerals (Natarajan, 1996). However, whether nitrogen purging
during grinding with iron grinding media can improve chalcopyrite flotation needs to be
confirmed.
Although mild steel grinding medium is detrimental to chalcopyrite flotation, it is
favoured for chalcopyrite selectivity against pyrite. Yuan et al. (1996) found that
chalcopyrite selectivity against pyrite was reduced when the minerals were ground in an
oxidising environment with a stainless steel medium, but was restored when ground in a
reducing environment with mild steel medium. A similar observation was found by Van
Deventer et al. (1991). Clearly, it is desirable to investigate the effect of grinding with low
iron content media on chalcopyrite separation from pyrite. Yuan et al. (1996) also
indicated that air-sparged mild steel grinding gave no advantage in selectivity compared
with ordinary mild steel grinding. However, oxygenation during conditioning and flotation
is considered an important factor in the selectivity of flotation operations. As reported by
Houot and Duhamet (1990) and Kuopanportti et al. (2000), oxygenation had a favourable
effect on the separation of chalcopyrite from pyrite and the oxygen demand increased with
the proportion of pyrite in mineral systems.
In this study, a specially designed mill which allowed the control of chemical
conditions during grinding was used. The effects of two types of iron media, mild steel
and 30 wt.% chromium (with approximate 70 wt.% iron), and three types of purging gases
(nitrogen, air and oxygen) on chalcopyrite flotation and chalcopyrite separation from
pyrite were investigated.

2. Experimental

2.1. Materials and reagents

Chalcopyrite and pyrite samples were obtained from Willyama, Earth History Supplies,
NSW, Australia and Huanzala Mine, Peru, respectively. The samples were crushed through
a rolls crusher and then screened to collect the + 0.6  3.2 mm particle size fraction. The
processed samples were then sealed in polyethylene bags. The chemical composition of
these samples analysed by ICP-MS is shown in Table 1.
Two types of grinding media were used. The tapered cylinder mild steel grinding
medium was supplied from Pasminco Mining, Elura, Australia and the spherical 30 wt.%
chromium grinding medium from Magotteaux, Australia. The specific surface areas of
Y. Peng et al. / Int. J. Miner. Process. 69 (2003) 87–100 89

Table 1
Chemical composition of the chalcopyrite and pyrite samples
Minerals Elements present (wt.%)
Si Fe Ca Mg Mn S Pb Cu Zn
Chalcopyrite 0.50 27.84 0.15 0.09 0.04 39.29 0.03 31.91 0.15
Pyrite 0.69 44.93 0.45 0.08 0.02 53.51 0.02 0.13 0.18

mild steel and chromium grinding media are 6.32  10 5 and 3.55  10 5 m2/g,
respectively, calculated by the measurements of the total surface areas of 4 kg of grinding
media.
Sodium diisobutyl dithiophosphate (DTP, AR grade) and Dowfroth 250 (polypropylene
oxide methanol, AR grade) were used in the experiments as collector and frother,
respectively. Copper nitrate (Cu(NO3)2, AR grade) were used to introduce copper species
during grinding. The pH was adjusted by the addition of AR grade NaOH solutions.
Deionised water was used in all experiments. High purity nitrogen and oxygen from BOC
Gases, Adelaide, Australia, were used throughout the study.

2.2. Mineral grinding and flotation

Chalcopyrite or pyrite single mineral (100 g) or chalcopyrite –pyrite mixture (50 g


each) was combined with 400 cm3 of water and ground with 4 kg of grinding medium in
the specialised mill for 30 min so that 90 wt.% of the particles were less than 53 Am in
diameter. The specific surface areas of mineral particles measured by BET after grinding
are shown in Table 2. The mill was originally designed by Cases et al. (1990). During
grinding, slurry was pumped through a monitoring cell where the pH was monitored and
fixed at 9.0 by adding NaOH solution continuously. The slurry in the monitoring cell was
purged with different gases to change the oxidation conditions in the mill. Samples for
EDTA extractions and XPS analyses were obtained from the cell at the completion of
grinding.
After grinding, the pulp was transferred to a 1.5-dm3 flotation cell, conditioned with pH
modifier, collector and frother. Four flotation concentrates were collected after cumulative
times of 0.5, 2.0, 4.0 and 8.0 min at an air flow rate of 2.5 dm3 min 1. The flotation froth
was scraped every 10 s. After grinding, the system was open to the air whilst the pH was
fixed at 9.0. In chalcopyrite – pyrite mixture mineral experiments, the chalcopyrite

Table 2
Specific surface areas of mineral particles after grinding
Grinding medium Mineral Specific surface
area (m2/g)
Mild steel Chalcopyrite 3.40
30 wt.% Chromium Chalcopyrite 3.39
Mild steel Pyrite 1.96
30 wt.% Chromium Pyrite 1.92
Mild steel Pyrite + Chalcopyrite 2.45
30 wt.% Chromium Pyrite + Chalcopyrite 2.04
90 Y. Peng et al. / Int. J. Miner. Process. 69 (2003) 87–100

recovery was calculated from the assay of copper, while pyrite recovery was calculated by
the difference in mass and chalcopyrite recovery.
The Eh (pulp potential) was measured at room temperature during grinding and
conditioning using a HANNA meter combined with a platinum electrode and expressed
relative to the standard hydrogen electrode, SHE.

2.3. Techniques

2.3.1. EDTA extraction


A 3 wt.% solution of AR grade ethylene diamine-tetra acetic acid disodium salt
(EDTA) was made up and a sodium hydroxide solution was used to adjust the pH to 7.5.
95 cm3 of the EDTA solution was placed in a vigorously stirred reaction vessel and purged
with nitrogen for 5 min. A 5 cm3 of slurry sample from the mill discharge was frozen in
liquid nitrogen first and then added to the EDTA solution, followed by 5 min of
conditioning. Nitrogen was continuously purged throughout. The slurry was then filtered
through a 0.45-Am Millipore filter. The filtrate was analysed for iron and copper by
inductively coupled plasma (ICP) atomic emission spectrophotometry (Spectroflame M).
The solids were retained to obtain the sample dry weight, enabling calculation of the mass
of metal oxidation species per unit mass of solid (Rumball and Richmond, 1996) and per
unit area of solid surface.

2.3.2. X-ray photoelectron spectroscopy (XPS)


The species on the surface of chalcopyrite after different grinding were examined by
XPS. XPS measurements were carried out with a Perkin-Elmer Physical Electronics
Division (PHI) 5100 spectrometer with a Mg Ka X-ray source operating at 300 W and
with a pass energy of 18 eV. The pressure in the analyzer chamber was 10 8 Torr during
analysis. The energy scale was calibrated using the Fermi edge and 3d5/2 line (BE = 367.9
eV) for silver, whilst the retardation voltage was calibrated using the position of the Cu
2p3/2 peak (BE = 932.67 eV) and the Cu 3p3/2 peak (BE = 75.13 eV). The measurements
were performed at a take-off angle of 45j. The slurry samples were washed with water of
pH 9.0 to remove any species from solution and introduced into the fore-vacuum of the
spectrometer as concentrated slurries. The samples were first examined in survey mode to
identify all the elements present, then the various elemental regions were scanned in order
to extract information on chemical bonding and oxidation stages. Details of this procedure
have been reported previously by Smart (1991).

3. Results and discussion

3.1. Effect of grinding conditions on chalcopyrite flotation

The effect of grinding media and gas purging on chalcopyrite flotation is shown in
Fig. 1. It can be seen that 30 wt.% chromium medium produced a higher chalcopyrite
recovery than mild steel medium. This is consistent with reports in literature that mild steel
medium depresses chalcopyrite flotation (Rao et al., 1976; Forssberg et al., 1988; Van
Y. Peng et al. / Int. J. Miner. Process. 69 (2003) 87–100 91

Fig. 1. Effect of grinding media and gas purging on chalcopyrite recovery as a function of flotation time: (dashed
lines) mild steel medium; (solid lines) 30 wt.% chromium medium; (5) nitrogen purging; (D) air purging; (o)
oxygen purging.

Deventer et al., 1991). Meanwhile, for both mild steel and 30 wt.% chromium grinding
media, gas purging had little effect on chalcopyrite flotation.
The samples from the mill discharge were analysed by EDTA extraction. The amount
of extracted iron and copper is shown in Table 3. The EDTA extracted copper originated
from the oxidation of chalcopyrite, whilst the EDTA extracted iron originated from the
oxidation of chalcopyrite and grinding media. According to the mixed-potential theory
(Fontana and Greene, 1978), when two materials are coupled, the oxidation rate of the
more active material will increase and the oxidation rate of the less active material will
decrease. It is well known that chalcopyrite is much more noble than the iron grinding
media (Rao et al., 1976; Kocabag and Smith, 1985). As a result, during grinding,
chalcopyrite oxidation is less pronounced than the medium oxidation and most EDTA
extracted iron should emanate from the grinding media. This is in agreement with the
results in Table 3 showing a much higher amount of EDTA extracted iron than copper.
From the amount of EDTA extracted copper, more chalcopyrite was oxidised after
grinding with 30 wt.% chromium medium than with mild steel medium. Oxygen purging
increased chalcopyrite oxidation as expected. Eh values at the completion of grinding are
also indicated in Table 3. A 30-wt.% chromium medium and oxygen purging resulted in
higher Eh values, suggesting more oxidising grinding conditions, which is consistent with
the higher amount of EDTA extracted copper.

Table 3
Effect of grinding conditions on the amount of EDTA extracted iron and copper in chalcopyrite single mineral
experiments
Grinding condition Eh Extracted Fe Extracted Cu
(medium; gas) (mV) (mol/m2 mineral) (mol/m2 mineral)
Mild steel; oxygen 195 1.19  10 4 1.07  10 6
Chromium; oxygen 220 3.26  10 5 1.42  10 6
Chromium; nitrogen 135 1.05  10 5 9.14  10 7
92 Y. Peng et al. / Int. J. Miner. Process. 69 (2003) 87–100

As discussed previously, iron oxidation species on chalcopyrite surfaces play the


dominant role in depressing chalcopyrite flotation. It can be seen that mild steel medium
produced a much higher amount of EDTA extracted iron than 30 wt.% chromium medium,
in agreement with the lower chalcopyrite recovery by mild steel medium. However,
oxygen purging also produced a higher amount of EDTA extracted iron than nitrogen
purging, but the flotation results in Fig. 1 were similar after grinding with oxygen or
nitrogen purging. Obviously, chalcopyrite flotation depends on the hydrophilic iron
oxidation species as well as some hydrophobic oxidation species on the chalcopyrite
surface.
The same samples for EDTA extraction were also analysed by XPS. The type of surface
species and their concentrations on chalcopyrite are reported in Table 4. As expected the
iron concentration on chalcopyrite surface is higher with mild steel medium than with 30
wt.% chromium medium, and with oxygen purging than with nitrogen purging. The Fe2p
spectra are shown in Fig. 2 and indicate that an iron oxyhydroxide layer is present on the
chalcopyrite surface. This is shown by a broad band in the Fe2p spectrum near 711– 712
eV and near 726 eV, while the iron associated with sulphur as in chalcopyrite occurs at 708
eV (Buckley and Woods, 1984). It can be seen that the relative intensity of the iron
oxyhydroxide band is higher when mild steel medium was used. However, this intensity
decreased after chalcopyrite was ground with 30 wt.% chromium medium, especially
under nitrogen purging. This observation is consistent with the amount of EDTA extracted
iron in Table 3.
The S2p spectra are shown in Fig. 3. According to Buckley and Woods (1984), the S2p
spectrum of chalcopyrite is composed of three components and each component is made
of a doublet. The peak near 161 eV is attributed to sulphide of unoxidised chalcopyrite.
The two higher binding energy components around 162 and 164 eV are due to metal-
deficient sulphide and polysulphide, respectively. Peaks for sulphate and thiosulfate at
higher binding energies around 168 eV (Buckley and Woods, 1984) are present when
oxygen was purged during grinding for both mild steel and 30 wt.% chromium grinding
media. Fig. 3 shows that 30 wt.% chromium medium produced more sulphate on the
chalcopyrite surface than mild steel medium, whilst oxygen purging produced more
sulphate and polysulphide than nitrogen purging in agreement with the EDTA exaction
results.
Although the type of gas used during grinding produced similar chalcopyrite recov-
eries, the type and proportion of the surface oxidation products were different. Higher
proportions of polysulphide, sulphate and ferric oxide/hydroxide were formed on
chalcopyrite surface with oxygen purging than with nitrogen purging. These results

Table 4
Concentration (at.%) of the elements measured by XPS on the surface of chalcopyrite
Grinding condition Atomic concentration of element (%)
(medium; gas) Fe S Cu O
Mild steel; oxygen 12.8 36.2 2.6 48.4
Chromium; oxygen 6.3 43.6 2.4 47.7
Chromium; nitrogen 4.7 58.4 2.3 34.6
Y. Peng et al. / Int. J. Miner. Process. 69 (2003) 87–100 93

Fig. 2. Fe2p XPS spectra of the chalcopyrite surface after grinding: (a) mild steel medium with oxygen purging;
(b) 30 wt.% chromium medium with oxygen purging; (c) 30 wt.% chromium with nitrogen purging.

indicate that chalcopyrite flotation is closely dependent on iron oxidation species, which
depress chalcopyrite flotation, and metal deficient sulphide, which increases chalcopyrite
flotation. Similar observations have been found by Fairthorne et al. (1997) who
demonstrated that the hydrophobicity of chalcopyrite will increase with the formation of
a sulphur-rich surface but will be offset by the precipitation of hydrophilic metal hydroxide
species.

3.2. Effect of grinding conditions on chalcopyrite separation from pyrite

The effect of grinding conditions on the flotation of chalcopyrite and pyrite in mineral
mixture experiments is indicated in Fig. 4. It can be seen that in agreement with the single
mineral flotation results (Fig. 1), 30 wt.% chromium medium produced higher chalcopyr-
94 Y. Peng et al. / Int. J. Miner. Process. 69 (2003) 87–100

Fig. 3. S2p XPS spectra of the chalcopyrite surface after grinding: (a) mild steel medium with oxygen purging; (b)
30 wt.% chromium medium with oxygen purging; (c) 30 wt.% chromium with nitrogen purging.

ite recovery than mild steel medium and gas purging had little effect on chalcopyrite
flotation. Furthermore, mild steel medium produced higher pyrite recovery than 30 wt.%
chromium medium with gas purging during grinding having no or little influence on pyrite
flotation. Therefore, higher chalcopyrite selectivity against pyrite was produced by 30
wt.% chromium medium than by mild steel medium.
Yelloji Rao and Natarajan (1989) reported that galvanic interactions between a noble
mineral and an active mineral markedly affected the floatability of the noble mineral, but
had a minimal effect on the floatability of the active mineral. The chalcopyrite flotation
rate constant and maximum recovery in the experiments of chalcopyrite single mineral and
chalcopyrite – pyrite mixture were calculated by fitting the experimental recovery data to
the first order rate equation (Lynch et al., 1981):
Recovery ðtÞ ¼ Rmax ð1  ekt Þ
Y. Peng et al. / Int. J. Miner. Process. 69 (2003) 87–100 95

Fig. 4. Effect of grinding media and gas purging on the separation of chalcopyrite from pyrite: (empty symbols)
mild steel medium; (solid symbols) 30 wt.% chromium medium; (squares) nitrogen purging; (triangles) air
purging; (circles) oxygen purging.

where Rmax (%) is the maximum recovery at infinite time and k (min 1) is the first order
rate constant. The rate constant and maximum recovery are shown in Table 5. After pyrite
addition, the chalcopyrite flotation rate constants decreased, but the maximum recoveries
remained within experimental error. It was also noted that pyrite exhibited a different
flotation behaviour after mixed with chalcopyrite. As shown in Fig. 5, pyrite single
mineral displayed a poor floatability with less than 10% recovery after 8 min of flotation.
However, pyrite flotation was increased after addition of chalcopyrite, especially with the
mild steel medium (Fig. 4).
Samples of chalcopyrite– pyrite mixture ground in different conditions were condi-
tioned with EDTA solutions. The extracted amounts of iron and copper are shown in Table
6. Higher iron oxidation production was produced with mild steel medium than with 30
wt.% chromium medium and with oxygen purging than with nitrogen purging. A
comparison of the amount of extracted iron in Tables 6 and 3 reveals similar trends and
only a small increase in iron oxidation after chalcopyrite was mixed with pyrite. This
increase in surface iron oxidation may contribute to the lower chalcopyrite flotation rate
constant observed in chalcopyrite – pyrite mixture experiments.
Attention now turns to the activation of pyrite flotation by chalcopyrite. Two different
mechanisms for the increased floatability of pyrite after chalcopyrite addition have been
proposed (Ekmekcß i and Demirel, 1997): (i) increased proportion of elemental or metal-

Table 5
Chalcopyrite flotation rate constant and maximum recovery in single mineral and mixture mineral experiments
Grinding condition Rate constant (min 1) Recovery maximum (%)
(medium; gas)
Cp Cp + Py Cp Cp + Py
Mild steel; oxygen 1.6 1.1 75 75
Mild steel; nitrogen 1.2 1.1 76 77
Chromium; oxygen 2.1 1.6 87 86
Chromium; nitrogen 2.4 1.9 88 86
96 Y. Peng et al. / Int. J. Miner. Process. 69 (2003) 87–100

Fig. 5. Pyrite single mineral flotation as a function of flotation time, with and without Cu2 + addition, under
.
oxygen purging: (o) without Cu2 +; ( ) with 1.5  10 3 M Cu2 +; (dashed lines) mild steel; (solid lines) 30 wt.%
chromium medium.

deficient sulphur on the pyrite surface and (ii) copper activation. Regarding mechanism (i),
because chalcopyrite is more active than pyrite, ferric hydroxide may be reduced to a more
soluble iron species, ferrous hydroxide, and thus elemental or metal-deficient sulphur will
be generated on pyrite surface, which results in an improved pyrite flotation in the
presence of chalcopyrite. A similar mechanism should also occur for pyrite in the absence
of chalcopyrite in this study as both the mild steel and 30 wt.% chromium grinding media
are more active than pyrite. However, Fig. 5 shows that pyrite flotation was poor in
absence of copper. This concluded that the improved pyrite flotation in the experiments of
chalcopyrite – pyrite mixture is not due to elemental or metal-deficient sulphur but rather to
copper activation of pyrite flotation. It is generally accepted that the formation of a new
copper sulphide phase on the pyrite surface through the reduction of copper(II) to
copper(I) is responsible for the activation of pyrite while sulphide ions are oxidised
(Bushell and Krauss, 1962; Weisener and Gerson, 2000). Laajalehto et al. (1999) proposed
that a chalcopyrite-like phase (CuFeS2) might be formed during the activation of pyrite
because the copper-activated pyrite exhibited the typical S2p XPS spectrum of copper
sulphides (e.g. chalcocite or chalcopyrite) and similar electrochemically controlled
xanthate adsorption behaviour as chalcopyrite. A comparison of the amount of EDTA
extracted copper in Tables 6 and 3 shows that a much lower amount of copper hydroxide

Table 6
Effect of grinding conditions on the amount of EDTA extracted iron and copper in chalcopyrite – pyrite mixture
experiments
Grinding condition Extracted Fe Extracted Cu
(medium; gas) (mol/m2 mineral) (mol/m2 mineral)
Mild steel; oxygen 1.25  10 4 2.57  10 7
Chromium; oxygen 4.43  10 5 5.78  10 7
Chromium; nitrogen 1.58  10 5 2.86  10 7
Y. Peng et al. / Int. J. Miner. Process. 69 (2003) 87–100 97

Fig. 6. Eh during grinding of pyrite in the presence of chalcopyrite (D) or Cu(II) (5) under oxygen purging:
(dashed lines) mild steel; (solid lines) 30 wt.% chromium medium.

was extracted from the chalcopyrite – pyrite mixture than from chalcopyrite in the absence
of pyrite even if considering that the amount of chalcopyrite in the mixture was half that in
single mineral. Therefore, it appears that the improvement in pyrite flotation in the
presence of chalcopyrite is due to copper activation. This was further confirmed by pyrite
flotation in the presence of copper(II). In Fig. 5, 1.5  10 3 mol/l of copper(II), based on
the amount of EDTA extracted copper in Table 3, was added during pyrite grinding. It can
be seen that addition of copper(II) had exactly the same effect on pyrite flotation as
chalcopyrite. Pyrite flotation was improved in the presence of copper(II), especially after
grinding with mild steel.
Fig. 6 shows the Eh during grinding as a function of grinding time. In the presence of
chalcopyrite or copper(II), mild steel medium always produced lower Eh values than 30
wt.% chromium medium. Voltammetric studies have shown that activation of pyrite by
copper is strongly dependent on the redox potential in the activating solution and that
increases in oxidising potential inhibit copper uptake (Richardson et al., 1996). This may
explain why pyrite flotation was increased further after grinding with mild steel medium in
the presence of copper(II) in Fig. 5. Since chalcopyrite displayed the same effect on pyrite
flotation as copper(II), the effect of grinding media on copper(II) reduction could be
studied by analysing the pyrite surface after grinding in the presence of copper(II) ions.
Table 7 shows that 30 wt.% chromium medium resulted in slightly more copper on the
pyrite surface than mild steel medium. However, it is copper(I) which is responsible for
pyrite activation. In order to quantify the amount of copper(I) on the pyrite surface, Cu2p

Table 7
Concentration (at.%) of the elements measured by XPS on pyrite surface in the presence of 1.5  10 3 M Cu(II)
Grinding condition Atomic concentration of element (%)
(medium; gas) Fe S Cu O
Mild steel; oxygen 20.8 20.7 1.7 56.8
Chromium; oxygen 20.3 26.4 2.1 51.2
98 Y. Peng et al. / Int. J. Miner. Process. 69 (2003) 87–100

Fig. 7. Cu2p XPS spectra of the pyrite surface in the presence of 1.5  10 3 M Cu2 +: (a) mild steel medium with
oxygen purging; (b) 30 wt.% chromium medium with oxygen purging.

XPS spectra were fitted as shown in Fig. 7. The relatively smaller intensity of the broad
band observed around 943.4 eV is attributed to copper(II) shake-up satellites and the peak
of 934.4 eV, attributed to cupric oxide or hydroxide indicates that the surface copper, is
mainly in a cuprous state with the most intense band at around 932.3 eV (Fairthorne et al.,
1997). The percentage of copper(I) on pyrite surface after grinding with mild steel medium
and 30 wt.% chromium medium is 79% and 54%, respectively, indicating that more pyrite
activation occurred with mild steel medium than with 30 wt.% chromium medium in
agreement with the flotation results in Fig. 5.

4. Conclusion

Grinding conditions had a significant effect on chalcopyrite flotation. This effect was
closely associated with the presence of iron oxidation species and metal deficient sulphide
present on the chalcopyrite surface. Iron oxidation species from grinding media played a
dominant role in depressing chalcopyrite flotation, whilst the presence of metal deficient
sulphide improved chalcopyrite flotation. As a result, 30 wt.% chromium medium
produced a higher chalcopyrite recovery than mild steel medium, whilst gas purging
during grinding had little effect on chalcopyrite flotation for the same grinding medium.
When chalcopyrite was mixed with pyrite, the amount of iron oxidation species from
grinding media was increased, which resulted in a lower rate of chalcopyrite flotation.
Y. Peng et al. / Int. J. Miner. Process. 69 (2003) 87–100 99

Pyrite flotation was significantly improved especially after grinding with mild steel
medium. This improvement in pyrite flotation was due to copper(II) activation. Grinding
media had a large effect on copper(II) reduction on pyrite surface. Mild steel medium
caused more reducing grinding condition and less pyrite oxidation than 30 wt.% chromium
medium. Thus, more copper(II) species were reduced to copper(I) species on the pyrite
surface, which resulted in more pyrite activation by mild steel medium. Consequently, 30
wt.% chromium medium produced better chalcopyrite selectivity against pyrite than mild
steel medium.

Acknowledgements

Financial support for this work from AMIRA International as well as scholarships to Y.
Peng from the University of South Australia are gratefully acknowledged. Pasminco
Mining, Elura, Australia and Magotteaux, Australia are thanked for the supply of grinding
media.

References

Buckley, A.N., Woods, R., 1984. An X-ray photoelectron spectroscopic study of the oxidation of chalcopyrite.
Aust. J. Chem. 37, 2403 – 2413.
Bushell, C.H.G., Krauss, C.J., 1962. Copper activation of pyrite. Can. Min. Metall. J., Trans. 65, 185 – 189.
Cases, J.M., Kongolo, M., deDonato, P., Michot, L., Erre, R., 1990. Interaction between finely ground galena and
potassium amylxanthate in flotation: I. Influence of alkaline grinding. Int. J. Miner. Process. 28, 313 – 337.
Ekmekcßi, Z., Demirel, H., 1997. Effects of galvanic interaction on collectorless flotation behaviour of chalco-
pyrite and pyrite. Int. J. Miner. Process. 53, 31 – 48.
Fairthorne, G., Fornasiero, D., Ralston, J., 1997. Effect of oxidation on collectorless flotation of chalcopyrite. Int.
J. Miner. Process. 49, 31 – 48.
Fontana, M.G., Greene, N.D., 1978. Materials Science and Engineering Series. McGraw-Hill, USA.
Forssberg, E., Sundberg, S., Zhai, H., 1988. Influence of different grinding methods on floatability. Int. J. Miner.
Process. 22, 183 – 192.
Houot, R., Duhamet, D., 1990. Importance of oxygenation of pulps in the flotation of sulphide ores. Int. J. Miner.
Process. 29, 77 – 87.
Kocabag, D., Smith, M.R., 1985. The effect of grinding media and galvanic interaction on the flotation of
sulfide mineral. Proc. Conf., Complex Sulfides-Proc. of Ores, Conc. and Byproducts, San Diego, CA,
U.S.A., pp. 13 – 20.
Kuopanportti, H., Suorsa, T., Dahl, O., Niinimäki, J., 2000. A model of conditioning in the flotation of a mixture
of pyrite and chalcopyrite ores. Int. J. Miner. Process. 59, 327 – 338.
Laajalehto, K., Leppinen, J., Kartio, I., Laiho, T., 1999. XPS and FTIR study of the influence of electrode
potential on activation of pyrite by copper or lead. Colloids Surf., A Physicochem. Eng. Asp. 154, 193 – 199.
Lynch, A.J., Johnson, N.W., Manlapig, E.V., Thorne, C.G., 1981. Mathematical models of flotation. Mineral and
Coal Flotation Circuits, Their Simulation and Control, Developments in Mineral Processing. Elsevier, Am-
sterdam, pp. 57 – 96.
Natarajan, K.A., 1996. Laboratory studies on ball wear in the grinding of a chalcopyrite ore. Int. J. Miner.
Process. 46, 205 – 213.
Rao, S.R., Moon, K.S., Leja, J., 1976. Effect of grinding media on the surface reactions and flotation of heavy
metal sulphides. In: Fuerstenau, M.C. (Ed.), Flotation, A.M. Gaudin Memorial Volume. SME-AIME,
New York, pp. 509 – 527.
100 Y. Peng et al. / Int. J. Miner. Process. 69 (2003) 87–100

Richardson, P.E., Chen, Z., Tao, D.P., Yoon, R.H., 1996. Electrochemical control of pyrite activation by copper.
In: Woods, R., Doyle, F.M., Richardson, P. (Eds.), Electrochemistry in Mineral and Metal Processing, vol. IV.
Pennington, New Jersey, pp. 179 – 190.
Rumball, J.A., Richmond, G.D., 1996. Measurement of oxidation in a base metal flotation circuit by selective
leaching with EDTA. Int. J. Miner. Process. 48, 1 – 20.
Smart, R.St.C., 1991. Surface layers in base metal sulphide flotation. Min. Eng. 4, 891 – 909.
Van Deventer, J.S.J., Ross, V.E., Dunne, R.C., 1991. The effect of milling environment on the selective flotation
of chalcopyrite from a complex sulphide ore. Proc. XVII Int. Min. Proc. Cong. Dresden, pp. 129 – 140.
Weisener, C., Gerson, A., 2000. Cu(II) adsorption mechanism on pyrite: an XAFS and XPS study. Surf. Interface
Anal. 30, 454 – 458.
Yelloji Rao, M.K., Natarajan, K.A., 1988. Influence of galvanic interactions between chalcopyrite and some
metallic materials on flotation. Miner. Eng. 1, 281 – 294.
Yelloji Rao, M.K., Natarajan, K.A., 1989. Electrochemical effects of mineral – mineral interactions on the flota-
tion of chalcopyrite and sphalerite. Int. J. Miner. Process. 27, 279 – 293.
Yuan, X.M., Pålsson, B.I., Forssberg, K.E., 1996. Flotation of a complex sulphide ore: II. Influence of grinding
environments on Cu/Fe sulphide selectivity pulp chemistry. Int. J. Miner. Process. 46, 181 – 204.

You might also like