Sie sind auf Seite 1von 11

Minerals Engineering 18 (2005) 855–865

This article is also available online at:


www.elsevier.com/locate/mineng

A review of pyrrhotite flotation chemistry in the processing


of PGM ores
a,*
J.D. Miller , J. Li a, J.C. Davidtz b, F. Vos b

a
Department of Metallurgical Engineering, University of Utah, Salt Lake City, UT 84112, USA
b
Department of Materials Science and Metallurgical Engineering, University of Pretoria, Pretoria 0002, South Africa

Received 20 January 2005; accepted 26 February 2005

Abstract

The chemistry of pyrrhotite flotation using xanthate collectors is reviewed with respect to the processing of PGM ores and the
recent results from captive bubble contact angle measurements at the University of Utah are presented. In some cases a low flotation
recovery of PGM may be due to the surface state of pyrrhotite particles under conventional flotation conditions (open to air and pH
9.0).
Thermodynamically pyrrhotite is not stable and reacts relatively quickly with its environment. Natural/collectorless flotation of
pyrrhotite is observed only under a low oxidation potential in acidic solution. Its surface is easily oxidized to ferric hydroxide/oxide
under conventional flotation conditions, creating a hydrophilic state at the pyrrhotite surface and low flotation recovery even
though xanthate collectors can be adsorbed. Under these conditions, activation by copper is not easily achieved. These observations
reported in the literature have been confirmed by captive bubble contact angle measurements. Based on the analysis of previous
research, conditions for improved pyrrhotite flotation and increased PGM recovery are suggested.
 2005 Published by Elsevier Ltd.

Keywords: Pyrrhotite; PGM; Flotation; Xanthate; Oxidation; Activation

1. Introduction the processing PGM ores from the Bushveld Complex in


South Africa.
Pyrrhotite is one of the most abundant iron sulfide The Bushveld complex is the worldÕs largest source of
minerals. It is found in nature to be commonly associ- platinum group minerals (PGMÕs). Within this deposit,
ated with pentlandite, quartz, ankerite (CaFe(CO3)2), two important horizons occur known as the Merensky
pyrite, chalcopyrite, and other sulfide minerals. In many reef and the UG2 reef which contain the valuable
flotation plants, pyrrhotite is rejected to the flotation PGM minerals. The PGMs are mainly associated with
tailings as a waste product such as in the case of Cu– base metal sulfides and to a lesser extent with oxides
Ni ores and massive nickel ores (Fornasiero et al., and silicate minerals. Typically the PGM ore contains
1995; Bozkurt et al., 1998; Chanturiya et al., in press; about 1% base metal sulfide minerals. The nature of
Khan and Kelebek, 2004). However, a strong interest the PGM association determines the overall flotation
in pyrrhotite recovery arises in certain instances. Specif- efficiency.
ically, the importance of pyrrhotite flotation is evident in There are three primary base metal sulfides in the
Merensky and UG2 reefs, namely pentlandite ((Fe,
*
Corresponding author. Tel.: +1 801 581 5160; fax: +1 801 581
Ni)9S8), chalcopyrite (CuFeS2), pyrrhotite (Fe1xS),
4937. with lesser amounts of pyrite (FeS2). The main sulfide
E-mail address: jdmiller@mines.utah.edu (J.D. Miller). mineral in the Merensky reef is found to be pyrrhotite,

0892-6875/$ - see front matter  2005 Published by Elsevier Ltd.


doi:10.1016/j.mineng.2005.02.011
856 J.D. Miller et al. / Minerals Engineering 18 (2005) 855–865

and consequently it is important to achieve high pyrrho- hydroxide (Smart et al., 2003; Buckley and Woods,
tite recoveries during flotation. 1985; Heyes and Trahar, 1984):
In general, under conventional flotation conditions
Fe1x S þ ð7  3xÞH2 O ¼ ð1  xÞFeðOHÞ3 þ SO2
4
open to air at pH 9.0, good chalcopyrite and pentlandite
recoveries are obtained with thiol collectors. However, þ ð11  3xÞHþ þ ð9 þ 3xÞe
pyrrhotite recoveries are not always satisfactory even ð3Þ
under conditions where xanthate adsorption is expected
Under these conditions, the natural flotation of pyrrho-
to take place (Buswell et al., 2002a). This may be due to
tite does not occur.
the fact that significant pyrrhotite oxidation occurs dur-
Recently at the University of Utah (2004), the natural
ing milling and the fact that pyrrhotite is a very poor
hydrophobicity of pyrrhotite (provided by the Geology
catalyst for oxygen reduction as may be required for
Curator, College of Mines and Earth Sciences, Univer-
the electrochemical adsorption of xanthate (Buswell
sity of Utah; Pyrrhotite content >95%, unknown source)
and Nicol, 2002b).
was examined in air from pH 3.0 to pH 9.2 based on
With more and more attention being given to pyrrho-
captive bubble contact angle measurements. The results
tite flotation in order to improve PGM recovery (Gathje
in Fig. 1 demonstrate that the pyrrhotite surface has a
and Simmons, 2004), a review of the flotation chemistry
strong hydrophilic state at pH values above pH 4.5 (con-
literature is presented together with some recent contact
tact angle 0). When the pH is less than pH 4.5, the nat-
angle measurements.
ural hydrophobicity of pyrrhotite increases with a
decrease in pH, having a contact angle of 51 at pH
3.0. These experimental results further confirm the flota-
2. Natural or collectorless flotation tion results reported by other investigators. It is evident
that a hydrophilic surface state is stabilized at pH >4.5
The natural floatability of pyrrhotite was investigated under an air atmosphere and room temperature.
by Hodgson and Agar (1984). These researchers sug- Further, pyrrhotite is thermodynamically unstable in
gested that a stable intermediate surface state, acidic solutions (pH  4.5) without consideration of oxi-
Fe(OH)S2, formed in acidic solution for short condition- dation. Theoretical solution chemistry calculation
ing times and at low oxidation potential of 0–200 mV regarding pyrrhotite stability by formation of H2S sug-
(SHE) accounting for the natural hydrophobicity ob- gests that H2S should evolve (1 atm total pressure) un-
served under certain conditions. However, extensive oxi- der an assumption of a ferrous ion activity of 104
dation during long conditioning times was found to lead (M) at pH 6 7 and determined by Fe(OH)2 solubility
to poor flotation. At the same time, it was reported by product at pH P 8. Thus, it is expected that such a reac-
Heyes and Trahar (1984) that pyrrhotite displays self- tion at the pyrrhotite surface may account for the natu-
induced flotation as a result of mild oxidation. Peters ral hydrophobicity in the absence of collector as
(1977) pointed out that pyrrhotite-collectorless flotation observed below pH 5 (see Fig. 1). Although the rate of
results from elemental sulfur formation at the mineral
surface since sulfur is strongly hydrophobic and may re-
main stable for a long time, even in alkaline solutions.
Similar research (Hamilton and Woods, 1981, 1983;
Heyes and Trahar, 1984; Buckley and Woods, 1985;
Jones et al., 1992) regarding the natural hydrophobicity
of pyrrhotite attributed its collectorless flotation to the
formation of iron-deficient/sulfur-rich metastable inter-
mediates at relatively low pH and under mild oxidation
potentials, which condition may be described by the fol-
lowing reactions:
Fe1x S þ 3yOH ¼ Fe1ðxþyÞ S þ yFeðOHÞ3 þ 3ye ð1Þ
and

Fe1x S þ 3ð1  xÞH2 O ¼ ð1  xÞFeðOHÞ3 þ S


þ 3ð1  xÞHþ þ 3ð1  xÞe
ð2Þ
However, pyrrhotite surfaces are oxidized rapidly upon Fig. 1. Contact angle at pyrrhotite surface as a function of pH at
exposure to air, and with an increase in oxidation time ambient temperature in 0.05 M Na2SO4 solution and in the absence of
the surfaces are covered by an overlay of iron(III) collector (University of Utah, 2004).
J.D. Miller et al. / Minerals Engineering 18 (2005) 855–865 857

this reaction has been reported to involve a significant 100


induction period, additional research is warranted. 90
pH 5.5
80

3. Flotation with thiocarbonate collectors pH 7.0


70

Flotation Recocery (%)


60
Thiocarbonates, particularly xanthates, are widely
used as collectors for selective and bulk flotation of sul- 50 pH 9.0
fide minerals from PGM ores. It has been found that 40
PGM recovery in bulk flotation is limited in some in-
stances by the poor flotation response of pyrrhotite. 30

With more and more attention being given to pyrrhotite 20


recovery, the use of thiocarbonate collectors has been
10
examined in order to identify conditions for increased
recovery. 0
0 1 2 3 4 5 6

3.1. Ethyl xanthate Ethyl Xanthate Concentration (M)

Fig. 3. Flotation recovery of pyrrhotite (conditioned 30 min) con-


Montalti (1994) systematically investigated pyrrhotite tacted with various concentrations of ethyl xanthate for 15 min at pH
floatability using ethyl xanthate. The pyrrhotite sample 5.5 (s), pH 7.0 (m), and pH 9 (h), [FeS] = 3.3 g/l in 2 · 103 mol/l
was from North Bend, Washington and was analyzed KNO3 (Montalti, 1994).
by XRD to be only one phase b-Fe1xS (Fe 47.0%
and S 32.6%) having hexagonal crystal structure with
some impurities (C 1.45% and Zn 4.64%). The zeta po- In this study by Montalti, the effects of pH, ethyl
tential of this pyrrhotite sample was measured under se- xanthate concentration, and collector reaction time, on
lected conditions and the results are presented in Fig. 2. flotation recovery were determined. It can be seen
The results show that the point of zero charge (PZC) from Fig. 3 that the influence of pH and collector con-
for this pyrrhotite mineral (sample 2 g/l, KNO3 centration is significant. The flotation recovery at
2 · 103 mol/l) is around pH 6.5 and that the electroki- 1 · 104 mol/l ethyl xanthate significantly increases
netic response is not significantly affected by the pres- from 50% to 82% with a pH decrease from pH 9.0 to
ence of ethyl xanthate at low concentration. These pH 5.5. When the collector concentration is above
results might be expected if the surface is oxidized and 1 · 104 mol/l, the flotation recovery was found to be
Fe(OH)3 is stabilized at the pyrrhotite surface since insensitive to variations in concentration. For these
Fe(OH)3 has a PZC of about pH 6.5. experiments, the conditioning time was 30 min, and
the reaction time for collector adsorption was 15 min.
The influence of ethyl xanthate adsorption time on
flotation recovery for different ethyl xanthate concentra-
tions was examined and the flotation recovery was
found to reach a maximum at 15 min. Of course, the ex-
tent of recovery is dependent on pH. A maximum flota-
tion recovery of 90% at pH 5.5 was achieved with an
initial concentration of 5 · 104 mol/l ethyl xanthate.

3.2. Butyl xanthate

Sodium n-butyl xanthate (1 · 104 mol/l) was used to


investigate the flotation selectivity between pentlandite
and pyrrhotite at pH 9.0 to 9.5 under potential control
(Khan and Kelebek, 2004). The results are shown in
Fig. 4. The samples employed in these tests originated
from process streams consisting mainly of pyrrhotite–
pentlandite middlings processed in a nickel–copper
plant in the Sudbury region of Canada. The middlings
Fig. 2. Zeta potential–pH plot for pyrrhotite in the presence of differ-
were first reground in order to liberate pentlandite from
ent concentrations of ethyl xanthate of 0 mol/l (s), 5.0 · 104 mol/l
(m), 1.0 · 103 mol/l (h), [FeS] = 2.0 g/l in 2 · 103 mol/l KNO3 pyrrhotite using mild steel grinding media in the pH
(Montalti, 1994). range of 9.0–9.5 (nitrogen atmosphere). The case for
858 J.D. Miller et al. / Minerals Engineering 18 (2005) 855–865

80 Potential Controlled:
-0.095 to - 0.005 V/SHE
70
Pentlandite Recovery %

60

50

40 Potential Uncontrolled:
0.25 to 0.30 V/SHE
30

20

10

0
0 10 20 30 40 50 60 70 80
Pyrrhotite Recovery %

Fig. 4. Flotation selectivity between pentlandite and pyrrhotite at pH Fig. 5. Contact angle at pyrrhotite surface as a function of pH in the
9.0 to 9.5; potential controlled: 0.095 to 000.5 V (SHE), potential absence/presence of SIBX at ambient temperature and 0.05 M Na2SO4
uncontrolled: 0.25–0.3 V (SHE); sodium isobutyl xanthate 1 · 104 M (University of Utah).
(Khan and Kelebek, 2004).

Even though most researchers claim that mild oxida-


potential control refers to a condition whereby the po- tion and/or low oxidation potential favors natural and/
tential of the pulp was initially kept at relatively low lev- or collector flotation of pyrrhotite, it was reported (Rao
els (0.095 to 0.055 V/SHE) by sparging nitrogen and and Finch, 1991; Hodgson and Agar, 1991) that under a
then at about 0.005 to 0 V/SHE for subsequent stages. nitrogen atmosphere pyrrhotite flotation with addition
In the case of uncontrolled potential, the potential dur- of xanthate is not possible. In some cases, when xan-
ing flotation with air continuously increased up to a thate is not oxidized to dixanthogen, the collector
level 0.25–0.3 V as in usual practice. adsorption density and degree of hydrophobicity may
Fig. 4 demonstrates that under potential control the be reduced.
selective flotation of pentlandite from pyrrhotite oc-
curs with sodium n-butyl xanthate. For uncontrolled
potential less flotation selectivity was observed. These 4. Activation
results suggest that the relatively high potential (open
to air during conditioning) favors pyrrhotite flota- In order to improve recovery in the xanthate flotation
tion in the system studied. A similar phenomenon of sulfide minerals, copper, lead, silver, and other metal
was observed in AlekseevaÕs research (1965), in which ions can be used as activators. Chang et al. (1954) con-
the surface oxidation did not apparently depress ducted a detailed experimental investigation of the acti-
pyrrhotite flotation. In the presence of xanthate, using vation and flotation of pyrrhotite with xanthates and
high levels of aeration, flotation recoveries of pyrrho- copper sulfate at pH 5.1. It was found that copper sul-
tite were improved as compared to low levels of fate significantly improves pyrrhotite recovery. In these
aeration. experiments, the cupric ions directly exchange with fer-
The floatability of pyrrhotite has recently been rous ions at the pyrrhotite surface, the activated state
considered from contact angle measurements at the Uni- accounting for the improved flotation of pyrrhotite:
versity of Utah (2004) using sodium isobutyl xanthate
Cu2þ þ FeS ¼ CuS þ Fe2þ ð4Þ
(SIBX) as collector for solution pH values from pH
4.5 to 9.2. During these measurements, the system was The concentration of copper, iron, and sulfur during
open to air. Fig. 5 suggests that the addition of collector activation of pyrrhotite was measured and the results
improves the hydrophobicity when compared with that presented in Fig. 6 confirm the activation reaction given
in the absence of collector, but the improvement de- in Eq. (4). Buswell and Nicol (2002b) electrochemically
pends on the system pH. When the pH is above pH 5 measured the anodic decomposition rate of CuS at a
the hydrophobicity of the pyrrhotite surface is relatively pyrrhotite surface. When the potential is above 0.4 V
low (contact angle 30) and not very sensitive to vari- (SHE), the anodic rate was found to become significant.
ation in collector concentration. At pH 4.5, the contact The standard equilibrium potential for CuS oxidation
angle increases greatly from 30 to 84 with the addition was calculated to be 0.548 V (SHE). In the absence of
of collector (1 · 104 M). copper, one of the anodic reactions for pyrrhotite is
J.D. Miller et al. / Minerals Engineering 18 (2005) 855–865 859

Fig. 6. Concentration of Fe, Cu, and S in solution during the reaction Fig. 7. Contact angle at pyrrhotite surface with and without copper
of pyrrhotite with copper sulfate solution at pH 5.1 (Chang et al., activation as a function of pH in the presence of SIBX (1 · 104 M) at
1954). ambient temperature and 0.05 M Na2SO4 (University of Utah, 2004).

found at a much lower potential of about 0.1 V (SHE).


Based on these results it seems that copper stabilizes the
pyrrhotite surface.
In contrast, the role and effectiveness of copper ions
for the activation of pyrrhotite in alkaline solutions is
not fully understood, and the conclusions are controver-
sial. Fundamental research (Nicol, 1984) demonstrated
that copper activation is not possible since copper is
essentially insoluble above pH 8 and thus not available
for reaction at the pyrrhotite surface. Furthermore, pyr-
rhotite particles are likely to be well oxidized and cov-
ered with ferric hydroxide/oxide, which may inhibit
any reaction with the underlying mineral surface. In
contrast, other studies (Kelebek et al., 1996; Leppinen,
1990; Senior et al., 1995) have shown that pyrrhotite
recoveries are significantly improved in alkaline solu-
tions with the addition of copper sulfate. Copper was
Fig. 8. Contact angle at pyrrhotite surface with and without lead
detected on concentrate particles from actual flotation
activation as a function of pH in the presence of SIBX (1 · 104 M) at
circuits operating at pH 9 (Yoon et al., 1995), indicating ambient temperature and 0.05 M Na2SO4 (University of Utah, 2004).
that pyrrhotite was effectively activated.
Recent experiments for activation of pyrrhotite by
cupric ions (sulfate) and lead(II) ions (nitrate) at an ini- 5. Other factors
tial concentration of 1 · 104 m sodium isobutyl xan-
thate in 0.05 M Na2SO4 have been carried out with In an actual flotation system, it is expected that the
contact angle measurements at the University of Utah situation becomes much more complicated since pyrrho-
(2004), Figs. 7 and 8. These results show that when tite particles coexist and/or are in contact with other
the pH is greater than pH 7 the presence of these metal minerals/materials. Under the various conditions, a pyr-
ions does not influence the contact angle at the pyrrho- rhotite particle may act as an anode or cathode, which
tite surface. When the pH is reduced to pH 5, these me- may lead to hydrophobic or hydrophilic surface states.
tal ions significantly increase the contact angle, from 28 From a review of the literature, there are such other fac-
to 84 for copper and from 28 to 51 to 73 for lead, tors influencing pyrrhotite floatability, including:
depending on concentration. It should be noted that Pyrrhotite-grinding media: the contact of pyrrhotite
copper activation is insensitive to concentration varia- with mild steel when being ground was found to have
tions under the conditions studied and that the contact a deleterious effect on its floatability (Admas et al.,
angle is greater than that obtained with lead activation. 1984). The reduced floatability was attributed to the
860 J.D. Miller et al. / Minerals Engineering 18 (2005) 855–865

formation of an iron hydroxide coating on the pyrrho- 10% contained hexagonal pyrrhotite only, and 9% con-
tite particles, which is not only hydrophilic, but also tained monoclinic pyrrhotite only. The remaining sam-
inhibits xanthate adsorption. ples, 8% contained troilite, stoichiometric FeS.
Pyrrhotite–pyrite: when pyrrhotite particles are con-
tacted/associated with pyrite particles during flotation, 6.2. Stability of pyrrhotite
pyrrhotite recovery can be improved. This phenomenon
is attributed to a shift of oxygen reduction from sites on Thermodynamically, pyrrhotite is not stable in aque-
pyrrhotite to pyrite, which consequently reduces iron ous solutions under conventional flotation conditions
hydroxide formation on pyrrhotite and promotes its flo- (pH 9.0, open to the air atmosphere, and 25 C), which
tation (Nakazawa and Iwasaki, 1985, 1986). is evident from the Eh–pH diagram in Fig. 9 (Garrels
Pyrrhotite–chalcopyrite: similar to pyrite influence, and Christ, 1990). At low levels of dissolved sulfur and
chalcopyrite acts as the cathode and pyrrhotite acts as iron pyrrhotite is stable between pH 7 and pH 9 at a po-
the anode. The presence of chalcopyrite accelerates the tential less than 0.35 V (SHE). It is expected that pyr-
oxidation reaction of xanthate on pyrrhotite (i.e. effec- rhotite is oxidized finally to hematite and sulfate under
tive collector adsorption) and increases pyrrhotite float- conventional flotation conditions.
ability. At the same time, a lower floatability of Kinetically, Montalti (1994) examined the dissolved
chalcopyrite was observed either due to iron hydroxide iron concentration from pyrrhotite, chalcopyrite, and
formation on its surface (as a cathode for oxygen reduc- pyrite in a neutral solution (pH 6.9) with variation in
tion) or to a shift of increased xanthate adsorption by the oxidation potential as shown in Fig. 10. It is evident
pyrrhotite. An XPS study confirmed that the increased that the rate and extent of iron release from pyrrhotite is
floatability of pyrrhotite is not due to activation of cop- much greater than that from chalcopyrite and pyrite. It
per ions but rather the absence of iron hydroxide on the was found that the rate of pyrrhotite dissolution at
pyrrhotite surface (Cheng and Iwasaki, 1992, 1994). 200 mV (SHE) decreases with an increase in reaction
time and that the dissolution rate becomes quite slow
after 3 h. It is expected that a portion of the dissolved
6. Discussion iron precipitated from the solution as Fe(OH)3, covering
the sample surface, and contributing to the decrease in
6.1. Pyrrhotite characteristics dissolution rate.
Because of the variation of pyrrhotite in composition
Pyrrhotite, Fe1xS, is nonstoichiometric and of vari- and crystal structure, each pyrrhotite sample may have a
able composition with a density which varies from different nature, which is supported by the following re-
4.58 to 4.65. Accordingly, it has some unusual charac- sults. Khan and Kelebek (2004) and Buswell and Nicol
teristics. First, the amount of sulfur varies from 50 to (2002b) examined the kinetics of pyrrhotite decomposi-
55 atoms of sulfur per 50 atoms of iron. That is, the x tion by electrochemical measurements—cyclic voltam-
value varies from 0 to 0.2 (when x = 0, the pyrrhotite
mineral is called troilite). Secondly, it has two symme-
tries for its crystal structure: when pyrrhotite is relatively
low in sulfur or x is close to 0, the mineral structure is
hexagonal or prismatic; when pyrrhotite is high in sulfur
its structure is monoclinic. Thirdly, pyrrhotite magne-
tism is very low when x is 0 (hexagonal) at 20 C, only
having a magnetic susceptibility of 1 · 105 e.m.u./g.
When x is close to 0.2 (monoclinic), its magnetism in-
creases to 13.1 e.m.u./g. So, next to magnetite, the sulfur
rich monoclinic pyrrhotite is the most common mag-
netic mineral. The bonding difference between troilite
and monoclinic pyrrhotite mineral samples has recently
been discussed based on XPS measurements (Skinner
et al., 2004).
In nature, pyrrhotite minerals occur with significant
variation in structure and/or composition. Generally, a
mixture of hexagonal (relatively iron-rich, h-Po) and
monoclinic (relatively sulfur-rich, m-Po) phases is pre-
dominant according to the literature. Arnold (1967) Fig. 9. Stability regions of iron oxides and sulfides in water at 25 C,
made the analysis of 82 terrestrial pyrrhotite samples. 1 atm. with Sum S = 1 · 106 mol/l and Sum Fe = 1 · 106 mol/l
Seventy percents were a mixture of h-Po and m-Po, (Garrels and Christ, 1990).
J.D. Miller et al. / Minerals Engineering 18 (2005) 855–865 861

Fig. 12. Cyclic voltammograms for FeS electrode in deoxygenated


Na2B4O7 (0.05 M) solution at pH 9.3, a scan rate 20 mV/s (Buswell,
2002).
Fig. 10. Concentration of iron in solution (pH 6.9) as a function of
time for various sulfide minerals at several Eh: chalcopyrite Eh
+200 mV (s), pyrite +200 mV (n), pyrite Eh +0 mV (j), pyrite Eh mum current density at 0.75 V (SHE). This anodic
200 mV (d), pyrrhotite Eh 200 mV (h) (Montalti, 1994). process is considered to involve oxidation of pyrrhotite
to sulfate in addition to the formation of ferric
hydroxide:
metry. KhanÕs results for a Sudbury pyrrhotite, as pre-
sented in Fig. 11, demonstrate that when the potential Fe1x S þ ð7  3xÞH2 O ¼ ð1  xÞFeðOHÞ3 þ SO2
4
is below 0.5 V (SHE) the initial rate of pyrrhotite oxida- þ ð11  3xÞHþ þ ð9 þ 3xÞe
tion (anodic current density) is quite slow even though ð3Þ
there is a peak for the oxidation in the potential region
from 0.05 V to 0.18 V (SHE). The maximum current In contrast, the results from electrochemical measure-
density for this peak is 0.2 mA/cm2 at pH 9.2 and oxy- ments (pH 9.3 and 0.05 M Na2B4O7) by Buswell and Ni-
gen concentration of 0.1 mg/l in the absence of collector. col (2002b) show that pyrrhotite from an RSA mineral
This peak was attributed to the following anodic dealer is much more active than the Sudbury pyrrhotite
reaction: sample used by Khan and Kelebek (2004), see Fig. 12.
These results suggest that when the potential exceeds
Fe1x S þ 3yOH ¼ Fe1ðxþyÞ S þ yFeðOHÞ3 þ 3ye ð1Þ
0.4 V (SHE), the initial rate of pyrrhotite oxidation
When the positive potential scan reaches 0.5 V (SHE), a (anodic current density) becomes significant. The anodic
new anodic current peak appears and reaches its maxi- current density varies from 1.0 to 2.0 mA/cm2 at a po-
tential 0–0.2 V (SHE).

6.3. Flotation response

It is evident that pyrrhotite is susceptible to oxidation


in aqueous solution, an important factor which influ-
ences its flotation recovery since the oxidation products
at particle surfaces inevitably influences surface hydro-
phobicity. The kinetics of pyrrhotite oxidation and the
oxidation products at the surface depend on:

(i) surface area of pyrrhotite;


(ii) oxidation potential/oxygen pressure;
(iii) solution pH;
(iv) temperature;
(v) time.
Fig. 11. Cyclic voltammogram of FeS electrode at pH 9.2, a scan rate
10 mv/s, DO = 0.1 mg/l, X = 0.0 (Khan and Kelebek, 2004). As a consequence all these factors will influence the
DO = dissolved oxygen, X = sodium n-butyl xanthate. natural/collectorless flotation of pyrrhotite.
862 J.D. Miller et al. / Minerals Engineering 18 (2005) 855–865

As shown in Figs. 9–12, pyrrhotite tends to be ther- Bozkurt et al. (1998) examined the IR spectra of pyrrho-
modynamically and kinetically unstable. It is generally tite with isobutyl xanthate adsorption (1 · 105 M, 9.2)
accepted that during the flotation process in acidic solu- (see Fig. 13). The bands in the spectra are well defined at
tion under low oxidation potentials, such as sparging 1265, 1048, and 1026 cm1, which are characteristic of
nitrogen, the surface products of pyrrhotite oxidation coupled S–C–S and C–O–C stretching vibrations in dix-
are rich sulfur intermediates and/or elemental sulfur, anthogen. These results are similar to those reported by
which accounts for the naturally hydrophobic pyrrhotite Leppinen et al. (1989), Leppinen (1990), and Valli and
surface. In alkaline solutions and under relatively high Persson (1994). It appears from Fig. 13 that the adsorp-
oxidation potentials—open to the air atmosphere, the tion density of dixanthogen at the pyrrhotite surface is a
surface is covered by ferric oxides and/or ferric hydrox- little greater for pyrrhotite alone (A) than for the pyr-
ide, resulting in a hydrophilic surface state. rhotite in direct contact with pentlandite (C). Additional
It is generally thought that thiocarbonate collectors studies by Rao and Finch (1991) established that the to-
adsorb at sulfide mineral surface due to an electrochem- tal xanthate adsorption was slightly greater in the pres-
ical reaction involving the formation of a dimer such as ence of air than with nitrogen. At the same time, the
shown by the following half cell reaction for the amount of adsorbed dixanthogen was found to be greater
dixanthogen formation: than adsorbed xanthate in the presence of air. Impor-
2ROCðSÞS ¼ ROCðSÞS–SðSÞCOR þ 2e ð5Þ tantly, these researchers observed that dixanthogen was
not produced in the nitrogen environment. These results
The standard potential of the xanthate/dixanthogen pair may help explain the fact that at a relatively low potential
in reaction (5) was reported to be 0.128 V/SHE and the selective flotation of pentlandite from pyrrhotite is
0.06 V/SHE (Buswell et al., 2002a; Buswell and Nicol, possible, as presented in Fig. 4, or that a nitrogen
2002b; Winter and Woods, 1973). Of course, the stan- environment (very low potential) inhibits pyrrhotite flo-
dard potential depends on the length of the hydrocarbon atability with xanthates. Also noting a high level of aera-
chain of the collector molecule (Plessis, 2004). Such a tion or oxidation potential for improved pyrrhotite
reaction is expected at the pyrrhotite surface coupled flotation (Khan and Kelebek, 2004; Alekseeva, 1965), it
with reduction of oxygen: is expected that the oxide or hydroxide product layer
1=2 O2 þ H2 O þ 2e ¼ 2OH ð6Þ from pyrrhotite oxidation may have a high electric resis-
tance (average resistivity for natural pyrrhotite: 1 ·
For the pyrrhotite–xanthate system under practical flo- 105 X/m, average resistivity for hematite: 1 · 101 X/
tation conditions, the pyrrhotite surface can be oxidized m, Shuey, 1975) and/or that xanthate oxidation may have
to hydroxides and/or oxides which inhibit pyrrhotite flo- significant over-potentials on the surface of hydroxide–
tation. However, in some cases, the pyrrhotite was oxide products. The high oxidation potential might be
found to be floatable. Thus, Khan and Kelebek (2004) needed to oxidize xanthate to dixanthogen and would
suggested that xanthate can form thermodynamically favor flotation of oxidized pyrrhotite particles. In con-
stable Fe(II)/Fe(III) hydroxyl xanthates in alkaline solu-
tions at potentials lower than the stability domain of
dixanthogen (Harris and Finkelstein, 1975; Wang
et al., 1989). During the initial stage of adsorption, xan-
thate electrostatically interacts with ferrous ion on the
surface.
þ þ
Po½FeðOHÞ þ X ! Po½FeðOHÞ X ! Po½FeðOHÞX
ð7Þ
This species is considered to be metastable and has a
low level of hydrophobicity, but it is considered to be a
favorable site to accommodate more xanthate and/or
dixanthogen. More stable iron compounds can follow:
Po½FeðOHÞX þ OH ! Po½FeðOHÞ2 X þ e ð8Þ
Further adsorption of xanthate can lead to dixantho-
gen formation:
Po½FeðOHÞ2 X þ X ! Po½FeðOHÞ2 X–X þ e ð9Þ
Fig. 13. FTIR spectra of pyrrhotite at xanthate concentration of
Po½FeðOHÞ2 X–X þ X ! Po½FeðOHÞ2 X–X2  þ e 5 · 105 mol/l and pH 9.2 in the cases of single mineral (A), mixed
minerals (B, solution contact with pentlandite), and mixed minerals
ð10Þ (C, direct contact with pentlandite) (Bozkurt et al., 1998).
J.D. Miller et al. / Minerals Engineering 18 (2005) 855–865 863

trast, it was reported that pyrrhotite was effectively de- Activation: according to the literature, the effective-
pressed and selective flotation of pentlandite and chalco- ness of copper for pyrrhotite activation in alkaline solu-
pyrite from pyrrhotite was achieved (Khan and Kelebek, tions is controversial. Some researchers claim that
2004; Kelebek, 1993) as a result of selective oxidation of copper activation is not possible since copper ions are
pyrrhotite at relative high levels of potentials through essentially insoluble in alkaline solution and thus not
pre-aeration of slurries. available for reaction. Furthermore, pyrrhotite particles
Even though adsorption of xanthate and/or dixanth- are likely to be well oxidized and covered with ferric
ogen can occur on the oxidized pyrrhotite particles, the hydroxide/oxide, which may inhibit any activation reac-
hydrophobicity is not great, and a low flotation recovery tion with the underlying mineral surface.
for pyrrhotite is frequently found under the traditional Other studies have shown that pyrrhotite flotation
flotation conditions as used in the PGM industry. These recovery is significantly improved with addition of cop-
phenomena indicate that adsorbed collector covers only per sulfate in alkaline solutions, and that copper has
a part of the oxidized pyrrhotite particles, a state which been detected on concentrate particles from actual flota-
needs to be examined in future research. tion circuits. These results obviously indicate that pyr-
rhotite may be activated by copper ions.
6.4. Summary The recent results of contact angle measurement at
the University of Utah support the first observation.
It can be noted from review of the literature and re- Addition of copper ions or lead ions did not improve
cent results at the University of Utah that the pyrrhotite the hydrophobicity of pyrrhotite in neutral and alkaline
flotation system is very complicated, resulting from var- solution, although some positive effect was observed be-
iation in the pyrrhotite composition/structure, pyrrho- low pH 7.
tite and thiocarbonate oxidation, Eh/pH, minerals
contacted/associated with it, and activation reactions.
The following chemical factors appear to have a signif-
icant influence on pyrrhotite hydrophobicity and its 7. Conclusions
recovery by flotation.
Oxidation potential: it has been seen in the above sec- The flotation recovery of PGMs is dependent on
tions that pyrrhotite is thermodynamically and kineti- the pyrrhotite flotation recovery since PGMs are nat-
cally unstable and is easily oxidized to ferric urally associated with this sulfide mineral. Pyrrhotite
hydroxide/oxide, or intermediate compounds, depend- is not stable and is easily oxidized to ferric hydrox-
ing on the oxidation potential level and solution pH. ide/oxide under conventional flotation conditions,
If the flotation process is operated under conventional open to air at pH 9.0. The products from pyrrhotite
conditions, it is unavoidable that the product from pyr- oxidation cover the pyrrhotite particles, rendering
rhotite oxidation is ferric hydroxide/oxide precipitating the particle surface hydrophilic and causing a
at particle surfaces, rendering the pyrrhotite surface lower flotation recovery of pyrrhotite under most
hydrophilic and causing its low flotation recovery even circumstances.
though adsorption of xanthates and formation of di- Under low oxidation potential in acidic solutions,
xanthogen occurs. In this regard, appropriate control pyrrhotite displays a distinct natural floatability. In the
of the oxidation potential should be important during presence of xanthate, however, if the potential is too
flotation process for improved flotation recovery. First, low to oxidize xanthate to dixanthogen, pyrrhotite flota-
the oxidation potential/oxygen level during milling tion is inhibited. It is evident that appropriate control of
should be controlled to protect pyrrhotite from oxida- oxidation potential should be important during the flo-
tion. Second, in order to assure formation of dixantho- tation process.
gen, relatively high oxidation potential (0–0.2 V/SHE) Generally, activation of pyrrhotite by copper and
should be maintained during a short conditioning stage lead is significant in acidic solutions, but in some cases
prior to flotation. has been found to be negligible in neutral and alkaline
pH: the pH effect has been shown to be very impor- solutions.
tant as is evident from the results for pyrrhotite natural Based on this review of the literature, it seems that
floatability, collector flotation, and activation. These re- improved pyrrhotite flotation to increase PGM recovery
sults demonstrate that under the traditional flotation might be achieved in at least two ways:
conditions, such as at pH 9–9.5, no natural pyrrhotite
floatability is observed (contact angle is zero); pyrrhotite • Development of new collectors for improved pyrrho-
floatability with collector is low; and activators (copper tite hydrophobicity in air at pH 9.0.
and lead) lose their effectiveness in most cases, precipi- • Control of Eh/pH to minimize oxidation of pyrrho-
tating as hydroxides. Significantly improved pyrrhotite tite and still achieve collector adsorption with activa-
flotation can be achieved at pH < 5. tion as might be necessary.
864 J.D. Miller et al. / Minerals Engineering 18 (2005) 855–865

Research in both these areas is in progress. The use of Hamilton, I.C., Woods, R., 1983. An investigation of the deposition of
sulfur on the gold electrode. J. Electrochem. 13, 783.
trithiocarbonates for pyrrhotite flotation is one example Harris, P.J., Finkelstein, N.P., 1975. Interaction between sul-
of collector development as an alternative to the tradi- fide minerals and xanthates. I. The formation of monothiocar-
tional xanthate collectors. With regard to oxidation po- bonate at galena and pyrite surfaces. Int. J. Miner. Process. 2 (1),
tential control, the use of inert gas flotation for the 77.
processing of PGM ores has been demonstrated (Gathje Heyes, G.W., Trahar, W.J., 1984. The flotation of pyrite and
pyrrhotite in the absence of conventional collectors. In: Richard-
and Simmons, 2004) and is being evaluated further. son, P.R., Srinivasan, S.S., Woods, R. (Eds.), Proc. Int. Symp.
Electrochem. Miner. Metal Process. Electrochem. Soc., Penning-
ton, NJ, 84–10, p. 189.
Acknowledgments Hodgson, M., Agar, G.E., 1984. An electrochemical investigation
into the natural flotability of pyrrhotite. In: Richardson, P.R.,
Srinivasan, S.S., Woods, R. (Eds.), Proc. Int. Symp. Electrochem.
Support for the PGM flotation research program at Miner. Met. Process. Electrochem. Soc., Pennington, NJ, 84–10,
the University of Utah and the University of Pretoria p. 189.
by the National Science Foundation (NSN Grant Jones, C.F., Lecount, S., Smart, R.St.C., White, T., 1992. Composi-
INT—0352807) and industrial participants, Impala tional and structural alteration of pyrrhotite surfaces in solution:
Platinum, Newmont, BOC, and Bateman, is recognized XPS and XRS studies. Appl. Surf. Sci. 55, 65–85.
Kelebek, S., 1993. The effect of oxidation on the flotation behavior of
with appreciation. nickel–copper ores. In: Proc. of XVIII Int. Miner. Process. Congr.,
pp. 999–1005.
Kelebek, S., Wells, P.F., Fekete, S.O., 1996. Can. Metall. Quart. 35,
References 329.
Khan, A., Kelebek, S., 2004. Electrochemical aspects of pyrrhotite and
Admas, K., Natarajan, K.A., Iwasaki, I., 1984. Grinding media wear pentlandite in relation to their flotation with xanthate. Part I:
and its effect on the flotation of sulfide minerals. Int. J. Miner. Cyclic voltammetry and rest potential measurement. J. Appl.
Process 12, 39–54. Electrochem., pp. 1–8.
Alekseeva, R.K., 1965. Tsventy Metally (Non-Ferrous Met.) 6 (3), 19. Leppinen, J.O., Basilio, C.I., Yoon, R.H., 1989. In situ FTIR study of
Arnold, R.G., 1967. Range in composition and structure of 82 natural ethyl xanthate adsorption on sulphide minerals under conditions of
terrestrial pyrrhotites. Can. Miner. 9, 31–50. controlled potential. Int. J. Miner. Process. 26, 259–274.
Bozkurt, V., Xu, Z., Finch, J.A., 1998. Pentlandite/pyrrhotite inter- Leppinen, J.O., 1990. FTIR and flotation investigation of the
action and xanthate adsorption. Int. J. Miner. Process. 52, 203– adsorption of ethyl xanthate on activated and non-activated sulfide
214. minerals. Int. J. Miner. Process. 30, 245–263.
Buckley, A.N., Woods, R., 1985. X-ray photoelectron spectroscopy Montalti, M., 1994. Interaction of Ethyl Xanthate with Pyrite and
oxidized pyrrhotite surfaces II. Exposure to aqueous solutions. Pyrrhotite Minerals, Dissertation, the University of South Austra-
Appl. Surf. Sci. 20, 280. lia, February 1994.
Buswell, A.M., Bradshaw, D.J., Harris, P.J., Ekmekci, Z., 2002a. The Nakazawa, H., Iwasaki, J., 1985. Effect of pyrite–pyrrhotite contact on
use of electrochemical measurements in the flotation of a platinum their flotabilities. Miner. Metall. Process. 2 (4), 206–220.
group minerals (PGM) bearing ore. Miner. Eng. 15 (March), 395– Nakazawa, H., Iwasaki, J., 1986. Galvanic contact between nickel
404. arsenide and pyrrhotite and its effect on flotation. Int. J. Miner.
Buswell, A.M., Nicol, M.J., 2002b. Some aspects of the electro- Process. 18, 203–215.
chemistry of the flotation of pyrrhotite. J. Appl. Sci. 32, 1321– Nicol, M.J., 1984. In: Richardson, P.E., Srinivasan, S., Woods, R.
1329. (Eds.), International Symposium on Electrochemistry Mineral and
Chang, C.S., Cooke, S.R.B., Iwasaki, I., 1954. Flotation characteristics Metal Processing. Electrochemical Society, pp. 152–168.
of pyrrhotite with xanthate. Trans. AIME 199, 209. Peters, E., 1977. In: Rand, D.A.J., Welsh, B.J. (Eds.), Trends in
Chanturiya, V., Makarov, V., Forsling, W., Makarov, D., Tro- Electrochemistry. Plenum Press, NY, p. 267.
fimenoko, T., Kuznetson, V., in press. The effect of crystallochem- Plessis, R. du 2004. The Thiocarbonate Flotation Chemistry of
ical pecularities of nickel minerals on flotation copper–nickel ores. Auriferrous Pyrite, Dissertation, 2004, the University of Utah.
Int. J. Miner. Process. Rao, S.R., Finch, J.A., 1991. Adsorption of amyl xanthate at
Cheng, X., Iwasaki, I., 1992. Effect of chalcopyrite and pyrrhotite pyrrhotite in the presence of nitrogen and implication in flotation.
interaction on flotation separation. Miner. Metall. Process. 9 (2), Can. Metall. Quart. 30, 1.
73–79. Senior, G.D., Trahar, W.J., Guy, P.J., 1995. Int. J. Miner. Process. 43,
Cheng, X., Iwasaki, I., 1994. An electrochemical study on cathodic 209.
decomposition behavior of pyrrhotite in deoxygenated solutions. Shuey, R.T., 1975. Semiconducting ore minerals, Developments in
Miner. Metall. Process. 11 (2), 160–167. Economic Geology, 4. Elsevier Scientific Publishing Company.
Fornasiero, D., Montalti, M., Ralston, J., 1995. Kinetics of adsorption Skinner, W.M., Nesbitt, H.W., Pratt, A.R., 2004. XPS identification of
of ethyl xanthate on pyrrhotite: in situ UV and infrared spectro- bulk hole defects and itinerant Fe 3d electrons in natural troilite
scopic studies. J. Colloid Interfacial Sci. 172, 467–478. (FeS)1. Geochim. Cosmochim. Acta 68 (10), 2259–2263.
Garrels, R.M., Christ, C.L., 1990. Solutions, Minerals and Equilibria. Smart, R.S.C., Amarantidis, J., Skinner, W.W., Prestidge, C.A.,
Jones and Bartlett Publishers, Boston and London. Vanier, L.L., Grano, S.R., 2003. Surface analytical studies of
Gathje, J.C., Simmons, G.L., 2004. Flotation of platinum group metal oxidation and collector adsorption in sulfide mineral flotation. In:
ore materials, United States Patent, Patent No.: US 6,679,383 B2, Wandelt, K., Thurgate, S. (Eds.), Solid–Liquid Interfaces. Top.
Date of Patent: January 20. Appl. Phys., 85, pp. 3–60.
Hamilton, I.C., Woods, R., 1981. An investigation of surface oxidation Valli, M., Persson, I., 1994. Interaction between sulfide minerals and
of pyrite and pyrrhotite by linear potential sweep voltammetry. alkylxanthates: 8. A vibration and X-ray photoelectron spectro-
J. Electrochem. 118, 327. scopic study of the interaction between chalcopyrite, marcasite,
J.D. Miller et al. / Minerals Engineering 18 (2005) 855–865 865

pentlandite, pyrrhotite and troilite, and ethylxanthate and decyl- Winter, G., Woods, R., 1973. The relation of collector redox potential
xanthate ions in aqueous solution. Colloids Surf. 83, 207–217. to flotation efficiency: monothiocarbonates. Sep. Sci. 8 (2), 261.
Wang, X., Forssburg, E.K.S., Bolin, N.J., 1989. Thermodynamic Yoon, R.H., Basilio, C.I., Marticorena, M.A., Kerr, A.N., Stratton-
calculations on iron-containing sulfide mineral flotation systems, I. Crawley, R., 1995. A study of the pyrrhotite depression mechanism
The stability if iron xanthate. Int. J. Min. Process. 27, 1. by diethylenetriamine. Miner. Eng. 8 (7), 807–816.

Das könnte Ihnen auch gefallen