You are on page 1of 14

Advances in Colloid and Interface Science 145 (2009) 97110

Contents lists available at ScienceDirect

Advances in Colloid and Interface Science


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

A review of the fundamental studies of the copper activation mechanisms for


selective otation of the sulde minerals, sphalerite and pyrite
A.P. Chandra, A.R. Gerson
Applied Centre for Structural and Synchrotron Studies, University of South Australia, Adelaide, South Australia 5095, Australia

a r t i c l e

i n f o

Available online 9 September 2008


Keywords:
Flotation
Copper activation
Collector
Sphalerite
Pyrite

a b s t r a c t
A review of the considerable, but often contradictory, literature examining the specic surface reactions
associated with copper adsorption onto the common metal sulde minerals sphalerite, (Zn,Fe)S, and pyrite
(FeS2), and the effect of the co-location of the two minerals is presented. Copper activation, involving the
surface adsorption of copper species from solution onto mineral surfaces to activate the surface for
hydrophobic collector attachment, is an important step in the otation and separation of minerals in an ore.
Due to the complexity of metal sulde mineral containing systems this activation process and the emergence
of activation products on the mineral surfaces are not fully understood for most sulde minerals even after
decades of research.
Factors such as copper concentration, activation time, pH, surface charge, extent of pre-oxidation, water and
surface contaminants, pulp potential and galvanic interactions are important factors affecting copper
activation of sphalerite and pyrite. A high pH, the correct reagent concentration and activation time and a
short time delay between reagent additions is favourable for separation of sphalerite from pyrite. Sufcient
oxidation potential is also needed (through O2 conditioning) to maintain effective galvanic interactions
between sphalerite and pyrite. This ensures pyrite is sufciently depressed while sphalerite oats. Good
water quality with low concentrations of contaminant ions, such as Pb2+and Fe2+, is also needed to limit
inadvertent activation and otation of pyrite into zinc concentrates. Selectivity can further be increased and
reagent use minimised by opting for inert grinding and by carefully choosing selective pyrite depressants
such as sulfoxy or cyanide reagents. Studies that approximate plant conditions are essential for the
development of better separation techniques and methodologies.
Improved experimental approaches and surface sensitive techniques with high spatial resolution are needed
to precisely verify surface structures formed after copper activation. Sphalerite and pyrite surfaces are
characterised by varying amounts of steps and defects, and this heterogeneity suggests co-existence of more
than one coppersulde structure after activation.
2008 Elsevier B.V. All rights reserved.

Contents
1.
2.

3.

Introduction. . . . . . . . . . . . . . . . .
Activation of sphalerite . . . . . . . . . . . .
2.1.
Mechanisms of Cu(II) activation . . . . . .
2.2.
Reaction kinetics . . . . . . . . . . . .
2.3.
Mechanisms of Cu(OH)2 activation . . . . .
2.4.
Zeta potential and isoelectric point . . . . .
2.5.
Copper concentration and activation duration.
2.6.
The effect of sphalerite iron content on copper
2.7.
Sphalerite surface oxidation . . . . . . . .
2.8.
Lead and iron sphalerite activation . . . . .
Copper activation of pyrite . . . . . . . . . . .
3.1.
Unactivated otation . . . . . . . . . .
3.2.
Mechanisms of Cu(II) and Cu(OH)2 activation .
3.3.
Effectiveness of activation . . . . . . . .

. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
activation
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .

Corresponding author. Tel.: +61 8 8302 3044; fax: +61 8 8302 5545.
E-mail address: andrea.gerson@unisa.edu.au (A.R. Gerson).
0001-8686/$ see front matter 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.cis.2008.09.001

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

98
98
98
100
100
100
101
102
102
103
104
104
104
105

98

A.P. Chandra, A.R. Gerson / Advances in Colloid and Interface Science 145 (2009) 97110

4.
Mixed pyrite and sphalerite otation
5.
Summary . . . . . . . . . . .
Acknowledgments . . . . . . . . .
References . . . . . . . . . . . .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

1. Introduction
Flotation is an important and versatile mineral processing step
used to achieve selective separation of minerals and gangue. It utilises
the hydrophobic (aerophilic) nature of mineral surfaces and their
propensity to attach to rising air bubbles in a waterore pulp as the
basis for separation [1]. Metal sulde minerals, for which this process
was originally developed, are generally weakly polar in nature and
consequently most have a hydrophilic surface [2]. Hence, collector
molecules such as xanthates and dithiophosphates are normally used
to increase hydrophobicity [3]. Certain sulde minerals such as
sphalerite (ZnS) do not respond well to short chain thiol collectors,
due to the relative instability of zincxanthate and hence require the
use of activators to enhance the adsorption between collector
molecules and the sphalerite surface [2,4]. The cupric ion (Cu2+),
generally in the form of sulfate or nitrate, is the most widely used
activator. Other heavy metal ions such as lead, silver, cadmium,
mercury and Fe2+/Fe3+ can also activate the sphalerite surface, but are
either not used commercially or are present as impurities within the
sphalerite lattice or in process water [5,6].
Separation of sphalerite through copper activation becomes
problematical when other minerals within the pulp are inadvertently
activated along with the sphalerite. Pyrite (FeS2) is one such mineral
that responds to copper activation and can be oated together with
sphalerite [2]. Being the most abundant sulde mineral pyrite is
undesirably associated, and in most cases ne grained and intimately
intergrown, with minerals of economic value [7]. This gangue pyrite is
a cause of reduced concentrate grade and increased smelting costs, for
most minerals such as sphalerite, chalcopyrite and galena. Pyrite is
also a primary contributor towards the substantial environmental
problem of acid mine drainage resulting in acidication of natural
water systems. Mining industry treatment costs, in the US alone, are
over $1 million/day [8].
Sphalerite and pyrite frequently occur together in ore deposits
along with galena and copper containing ores such as chalcopyrite.
The common practice in mine otation is to rst oat the copper
containing minerals (if present) followed by galena [2,9]. The tails
from the galena otation are then used to oat sphalerite away from
pyrite primarily using copper activation. Effective separation is
needed to minimise iron in the nal zinc concentrate that may
occur through pyrite copper activation and otation. Loss of selectivity
and unwanted activation can also occur due to contaminants present
in the mine water used for separation of these minerals. Despite
continuous process improvements the problem of pyrite misreporting
to sphalerite concentrates still remains [10,11].
A review of the key factors affecting copper activation of sphalerite
and pyrite is presented herein. Special attention is given to the role of
sphalerite iron content. A discussion of the proposed activation
products as identied by various fundamental studies is provided as is
an examination of the literature regarding the activation and otation
response of the mixed sphalerite/pyrite system. While only sphalerite
and pyrite activation is discussed, the issues raised may also apply to
other sulde minerals.
2. Activation of sphalerite
The activation of sphalerite has been studied extensively over several
decades [5,1214]. While there is general agreement on the overall
process of copper activation, the actual mechanism and surface reaction
products controlling activation/otation still remains controversial.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

105
107
109
109

It has been well established that copper activation of sphalerite


follows an ion exchange mechanism where the uptake of Cu(II)1
results in approximately 1:1 release of Zn2+ into the solution [5,13,15]
and is generally represented by Eq. (1) [16].
2
ZnSs Cu2
aq CuSs Znaq

Cu(II) on the sphalerite surface is subsequently reduced to Cu(I) with


the resulting oxidation of the surface sulde. Collector molecules, such
as xanthates, then react with the surface copper sulde species formed,
thus increasing the otation response [17]. Cu(I)xanthate is the main
surface product formed, especially at low pH [15]. The uptake of copper
and subsequent otation of sphalerite is seen to depend on impurities,
such as iron, present within natural sphalerite, surface oxidation,
copper and xanthate concentration, activation time, solution O2
concentration (pulp potential) and most importantly on pH.
Under contrasting pH conditions either an abundance of hydrophobic (collector absent) or hydrophilic species maybe present on the
sphalerite surfaces [14]. Hydrophobic species such as polysuldes (S2
n )
and elemental sulfur (S0n) appear to predominate at mildly acidic
conditions while hydrophilic species such as zinc hydroxide and
copper hydroxide, along with some sulte/sulfate, occurs at higher pH
[14,15,18]. Polysuldes or elemental sulfur forms as a result of
oxidation of metal-decient sulde on sphalerite surfaces, however
it is still unclear if one or both of these products form and their relative
abundance on the surface [14,18,19]. The presence of such hydrophobic
species tends to promote collectorless otation of sphalerite which is
more prevalent at low pH [5,15]. Collectorless otation of sphalerite
may also be seen when impurities such as copper and iron, diffuses
from the bulk to the surface under acidic conditions [20]. This happens
after zinc dissolution when bulk cationic impurities migrate to the
metal-decient (sulfur-rich) sphalerite surface resulting in a selfactivating mechanism. Surface hydrophilic species are seen to reduce
recovery by otation with xanthate collectors, only within the pH
range where such species are stable [14].
Fig. 1 shows the various processes that may take place simultaneously during copper activation of sphalerite that result in the
production of hydrophobic and hydrophilic species. These processes
may also occur during activation of other suldes. Selective adsorption
of Cu(II) is the desired process. However, depending on activation
conditions, precipitation of hydrophilic copper containing hydroxide
species onto sphalerite surface takes place. In addition, aqueous Cu2+
and/or Cu(OH)2 also reacts with collector molecules leading to nonselective adsorption of hydrophobic CuX or dixanthogen (X2) on
sphalerite surfaces.
2.1. Mechanisms of Cu(II) activation
It has been noted that out of the many stable and metastable
intermediates present in a copper/sulfur system, researchers have
mainly considered CuS (covellite) and Cu2S (chalcocite) as possible
end products of copper activation of sphalerite (Buckley et al. [20]).
Buckley et al. [21] found a Cu 2p3/2 Auger parameter of 1850.0 eV on
copper-activated sphalerite which was about 0.3 eV lower than that
for covellite and hence suggest that even if a CuS-like phase forms it is
not same as covellite. They also did not nd any evidence of elemental
sulfur formation. Buckley et al. [20] describe the activated sphalerite
1
The nomenclature (I), (II) or (III) is used throughout to denote surface or bulk
species while superscript oxidation state numbers are used to denote aqueous species.

A.P. Chandra, A.R. Gerson / Advances in Colloid and Interface Science 145 (2009) 97110

Fig. 1. Schematics of sphalerite copper activation showing the various simultaneous


processes likely to occur under different activation conditions.

surface as a copper-substituted sphalerite lattice with the formation of


a metal-decient sulde (sulfur-rich) surface layer in both acidic and
alkaline media. This, according to them, is better represented by
Eq. (2), than by Eq. (1), where the product is metastable.
2
ZnSs xCu2
aq Zn1x Cux Ss xZnaq

Using conventional X-ray diffraction (XRD), scanning electron microscopy (SEM) coupled with energy dispersive spectrometry (EDS) and
electron microprobe analysis (EMPA), Vinals et al. [22] found digenite
(Cu1.8S) to be the main activation product at pH 1.11.3 and
temperatures of 180212 C while chalcocite appears to predominate
at 225 C. These temperatures are well above the temperatures found
within mineral processing plants. Chen and Yoon [23] conducted rest
potential measurements and voltammetry at pH 9.2 using a carbon
matrix composite (CMC) electrode containing sphalerite particles, to
show that when copper activation is carried out at open circuit
conditions a CuS-like activation product forms while activation
conducted at lower potentials produces a Cu2S-like product. They
further showed that activation conducted under slightly oxidising
conditions produced hydrophobic species, such as copper polysuldes, on the surface.
Pattrick et al. [17] used Cu K edge and S K edge X-ray absorption
spectroscopy (XAS, consisting of both extended X-ray absorption ne
structure, EXAFS, spectroscopy and X-ray absorption near edge
spectroscopy, XANES) data to show that copper on activated
sphalerite, at pH 1012, exists in a tetrahedrally coordinated form,
bonded to three sulfur atoms and one oxygen atom. However, these
measurements were carried out on dry samples and hence chemadsorbed water may have been present that is not localised on the
adsorbed copper atoms within wet slurry. Upon addition of xanthate
collector the oxygen of the CuO (2.07 ) bond is replaced by sulfur
from the xanthate and a primitive covellite species forms.
However, also using XAS, it was shown by Gerson et al. [13] that
under mildly acidic conditions both bulk and surface copper is
coordinated only to three sulfur atoms in a distorted trigonal planar
geometry with a CuS average bond length of 2.27 0.02 . These
measurements were carried out using wet slurry. It was proposed that
this geometry cannot be attributed to the formation of a distinct
crystallographic copper sulde phase and the data could not be
adequately tted to a structural model encompassing CuO bonds. It
has been further shown in Gerson et al. [24] that this distorted trigonal
planar structure (CuS3) on the sphalerite surface is slightly pushed up

99

and outwards resulting from the shorter CuS bonds as compared to


bulk ZnS bonds, and elongation of the attaching ZnS bonds. It was
found that the copper-activated sphalerite surface has CuS bonds
0.009 0.001 nm shorter than bulk ZnS bonds with a copper to zinc
distance 0.013 0.006 nm longer than the zinc to zinc distance.
Buckley et al. [21] recently examined a relatively pure copperactivated sphalerite surface (from Santander, Spain) using synchrotron
X-ray photoelectron spectroscopy (XPS, Cu, Zn and S 2p3/2 binding
energies) and XANES (Cu L2,3 edge). This study reconrmed that
copper on activated sphalerite surface exists predominantly as Cu(I),
although a higher than expected d9 character was found. Moreover it
was also found that some Cu(II) ions were present on the surface,
associated with oxygen (Cu(II)O species) possibly due to chemisorbed water on the sphalerite surface. Buckley et al. [21] suggested
the formation of a metastable phase (Zn1 xCuxS) with formal
oxidation state of sulfur being more positive than 2. This notion of
S being more positive is consistent with the model proposed by
Gerson et al. [13]. Furthermore, on the basis of the poor correlation
seen between the concentration of high binding energy sulfur and
surface copper concentrations, Buckley et al. [21] suggest unsubstituted (by copper) oxidative losses of zinc (Eq. (3)) from the sphalerite
lattice which leads to the development of enhanced sulfur regions
with oligosulde-like electronic environments.

Zn1x Cux Ss Zn1xy Cux Ss yZn2


aq 2ye

Kartio et al. [19] suggested that there may be difculties in assigning


Cu 2p3/2 binding energies (even from synchrotron XPS) to either CuS
or Cu2S. The variable XPS S 2p binding energies observed have led to
the suggestion that sulfur was present in different oxidation states
(non-integer) depending on the extent of copper activation and most
proposed mechanisms fail to account for this [13]. On this basis it was
proposed that increased copper activation resulted in less negative
sulfur oxidation states. The reaction mechanism provided for surface
substitution (Eqs. (4) and (5) for an initial and then a second copper
substitution respectively) and bulk substitution (Eq. (6) for a single
substitution) of zinc with copper result in the formation of a similar
distorted trigonal planar geometry with copper bonded to three sulfur
atoms but variable sulfur oxidation states. Note in Eq. (5) the less
negative oxidation state of the bridging sulfur atom.
2 2 4
0:9 1:63 4
Cu2
surface Zn2
aq Zn S3 surface Cu S3
aq

6
1:63
Cu2
Cu0:9 S1:63 Zn2 S2
2 surface
aq S2

2
S1:63
Cu0:9 S1:28 Cu0:9 S1:63
6
2
2
surface Znaq

2
2
2 0
2
1:63
Cu0:8 4:09
S2
3 Zn S Zn3 bulk Cuaq S3
bulk
2
S0:8 Zn1:63
6
3
bulk Znaq

Activation by Cu(II) results in the formation of a conducting layer


(with the band gap as low as 1.1 eV) on sphalerite surfaces [14,19].
Sphalerite is naturally an insulator with a band gap of 3.5 eV. This
reduced band gap aids in electron transfer reactions and allows the
thiol collectors to form an insoluble collector complex on the
sphalerite surface through mixed potential electrochemical reactions
[19,20]. Ab initio cluster model calculations have also shown that
copper atoms incorporated into the sphalerite lattice enhanced the
electron acceptor ability of sphalerite [25]. Cu(I)xanthate and
dixanthogen formation in solution in acidic medium is proposed to
occur via Eq. (7) [26]. Dixanthogen adsorption is more prevalent on
unactivated sphalerite [17] or when the Cu(II) concentration is low [4].

2Cu2
aq 4EXaq 2CuEX2aq 2CuEXaq X2aq

100

A.P. Chandra, A.R. Gerson / Advances in Colloid and Interface Science 145 (2009) 97110

2.2. Reaction kinetics


The initial stage of the activation process, generally, is very fast
with a high rate of copper uptake within rst 10 to 15 min [15,19,27].
This initial copper migration into the sphalerite surface can have a
diffusion coefcient of 1.0 10 11 to 1.5 10 10 cm2s 1 in the
temperature range of 80 to 400 C [28]. It has been proposed that
the rate of reaction is rst order with respect to copper concentration,
with the rate controlled by diffusion through the solution phase [5].
Thereafter the rate of copper uptake generally slows and shows
logarithmic time dependency [19] which is associated with copper
diffusion into the bulk of sphalerite lattice [5,29]. Vinals et al. [22]
found that this latter reaction rate followed a parabolic kinetic law at
high temperatures (160225 C) and had a high activation energy of
147 kJmol 1. This indicated a kinetic control of the reaction by solidstate counter diffusion of Cu(I) and Zn(II) ion through the copper
sulde layer.
The two step activation kinetics observed may be due to a
difference in mechanism for the adsorption/absorption of surface
and bulk copper. It has been proposed, based on EXAFS data, that
copper is incorporated onto the surface via replacement of three ZnS
bonds with commensurate low activation energy [13]. However, the
bulk copper results from replacement of four ZnS bonds to form only
three CuS bonds which requires greater activation energy.
XPS depth proling showed signicant Cu 2p signal even after
15 nm of argon ion etching [14,18]. Such copper diffusion processes
may however reduce the formation of surface Cu(I)xanthate
especially in dilute acidic solutions [26]. Popov and Vucinic [26]
showed, through IR internal reection measurements, that when the
activation time was long and copper was allowed to diffuse into the
bulk sphalerite structure, the concentration of Cu(I) xanthate on the
surface was reduced. Flotation studies however do not show any
signicant decrease in sphalerite recoveries as a result of this decrease
in surface xanthate (refer Section 2.5 for further details) [30].
2.3. Mechanisms of Cu(OH)2 activation
Prestidge et al. [18] studied the activation of synthetic sphalerite,
conditioned for 30 min at pH 9, using different concentrations of
copper nitrate. Using conventional XPS, they showed that at high
copper concentrations (10 or more monolayer coverage) the sphalerite surface becomes heavily coated with Cu(OH)2. SIMS results by
Gerson et al. [13] conrmed the occurrence of colloidal Cu(OH)2 on
sphalerite particles at high pH and high nominal copper coverage.
Similar results were also found by Fornasiero and Ralston [14], while
Popov and Vucinic [15] also noted an apparent depressing effect on
sphalerite otation at high copper concentration and alkaline pH.
Prestidge et al. [18] explain that Eq. (2) is more representative of
copper activation in an acidic medium where the activating species is
only Cu2+. According to them, activation in alkaline conditions where
surface Cu(OH)2 precipitate is found to be the predominant activating
species, is better represented by Eqs. (8), (9).
nZns xCuOH2ppt ZnSn d xCuOH2surface

The Cu(II) from the hydroxide may then exchange with the Zn(II)
from the sulde.
ZnSn d xCuOH2surface Znn xCux d xZnOH2surface

The zinc hydroxide formed undergoes dissolution and/or dispersion,


the extent of which controls the surface hydrophobicity [14,18]. The
resulting Cu(II) sulde then undergoes redox disproportionation to
form Cu(I) sulfur products. These products may then form Cu(I)
xanthate upon collector addition by combining with xanthate [17]. At
increased copper concentration the copper-substituted zinc sulde
layer becomes coated with an inhibiting copper hydroxide over-layer

[18]. A similar process has also been observed for pyrite [31] however
Cu(I) was observed on the pyrite surface prior to Cu(OH)2 deposition.
There has been suggestions that X-ray based techniques (especially
in ultrahigh vacuum, e.g. XPS) may cause an increase in the Cu(I) signal
from activated sphalerite, through photoreduction of Cu(II) to Cu(I)
[5,18]. Such a reduction process has now been shown to be conned
only to Cu(OH)2 over-layers, at extended X-ray irradiation time, and
does not affect Cu(II) involved with activation [32]. An associated
reduction in the surface concentration of oxygen was however
observed, and authors have suggested the use of a N2-cooled sample
holder to minimise this loss. An additional experimental uncertainty
may result from the necessity to conduct measurements on dry
samples where the surface species may not necessarily be representative of those in mineral slurry. The use of in situ techniques such as
IR will be ideal as it eliminates the need to house samples in ultrahigh
vacuum.
Some researchers believe that for sphalerite in alkaline media the
surface Cu(OH)2 directly interacts with the xanthate, where the OH
ion is exchanged with the xanthate ion [4,15]. The resulting product
then decomposes to form Cu(I)xanthate and dixanthogen on the
surface (Eqs. (10) and (11)).
CuOH2surface 2EXaq CuEX2surface 2OHaq

10

2CuEX2surface 2CuEXsurface EX2surface

11

2.4. Zeta potential and isoelectric point


The zeta potential and the isoelectric point (iep) of sphalerite are
also inuenced by the surface speciation resulting from variations in
pH, conditioning time, reagent, and reagent concentrations
[14,15,33,34]. Popov and Vucinic [15] used microelectrophoresis and
performed electrokinetic measurements on sphalerite particles,
conditioned with different reagents of varying concentration, as a
function of pH. It was found that sphalerite on its own had positive
zeta potential in acidic conditions and negative zeta potential in
alkaline conditions with the iep at pH 6.5. This remained the same
when sphalerite was conditioned with xanthate.
However conditioning sphalerite with copper sulfate solution
yielded a different zeta potentialpH curve with copper solutions of
different concentrations resulting in different sphalerite zeta potential
pH dependencies. Zeta potential for sphalerite was negative below pH
6 for all copper concentrations used, which according to Popov and
Vucinic [15] showed the exchange of Cu(II) with Zn(II) from
the sphalerite lattice. The negative zeta potential value suggests that
the surface was stabilised and no H+ adsorption took place. Zeta
potential of sphalerite conditioned with higher copper concentration
(8.0 10 4 moldm 3) showed two charge reversals with increasing
pH. According to Popov and Vucinic [15], the rst charge reversal from
negative to positive at pH 6 was due to adsorption and precipitation of
+
positive hydrolysed copper ion species (Cu2(OH)2+
2 , Cu(OH) ) predominantly on the sphalerite surface while the second charge reversal
from positive to negative at pH 7.6 resulted from deprotonation of the
copper hydroxide that predominantly covered the sphalerite surface.
This behaviour is in contrast to the less concentrated copper solution
(1.56 10 4 moldm 3) which showed a signicant decrease to more
negative zeta potential values from pH 6 to 8 and no charge reversals.
Sphalerite conditioned with copper and xanthate showed negative
zeta potential throughout the pH range (5.8 to 9.2) due most likely to
the adsorption of xanthate (X) onto the sphalerite surface. Charge
reversals are usually common in mineral oxides, silicates and suldes
in the presence of adsorbing metal ions and this behaviour can vary
depending on type of metal ion, its salt and its concentration [35].
Zhang et al. [34] used a similar technique to study the zeta
potential of sphalerite alone and conditioned with various concentrations of ferrous ion and xanthate as a function of pH. While basic zeta

A.P. Chandra, A.R. Gerson / Advances in Colloid and Interface Science 145 (2009) 97110

potentialpH trends obtained by Zhang et al. [34] were similar to


those from Popov and Vucinic [15], the sphalerite zeta potential values
of Zhang et al. [34] were predominantly negative for all conditions
studied throughout the pH range (212) and the iep of sphalerite was
located at pH 2.5. A signicant difference in these two studies was the
iron content of sphalerite samples used. Popov and Vucinic [15] used a
natural sphalerite sample containing approximately 13 wt.% iron
while Zhang et al. [34] also used a natural sphalerite sample but
containing only 2.8 wt.% iron.
Sphalerite samples with different iron contents are known to
exhibit different zeta potential values [33]. Gigowski et al. [33]
demonstrated that the zeta potential and iep of sphalerite varies nonlinearly with the iron content of sphalerite. Mirnezami et al. [36] also
suggests the likelihood of iron content inuencing iep (and hence zeta
potential) of sphalerite. While investigating the aggregation mechanism of sphalerite particles Mirnezami et al. [36] measured the
electrophoretic mobilities of sphalerite of different iron contents
and found that the iep of the sample with low iron (mineral sample)
was at a lower pH than the samples with higher iron content (plant
concentrates). This was attributed to the presence of increased
concentrations of zinc oxidation products on the plant concentrates,
however the added inuence of iron was also suggested. Sphalerite
surface charge thus appears to be an important factor to be considered
in the differential activation/otation response of sphalerite with
different iron contents (refer Section 2.6), normally observed in
fundamental studies.
In a recent study, Fornasiero and Ralston, [14] measured electrophoretic mobilities of sphalerite, conditioned at either pH 6.0 or 8.5,
with increasing copper concentrations and used XPS measurements to
identify the surface species causing the observed change in zeta
potential with pH. The zeta potential trend with pH was similar to
those found by Popov and Vucinic [15] and Zhang et al. [34]. At pH 6.0,
polysuldes were one of the species found on the surface of sphalerite
responsible for decreasing the zeta potential to negative values. These
polysuldes may have appeared as a result of the formation of metaldecient sulfur-rich phases formed on metal dissolution and may
possibly exist as (Cu+)2S6 [19] or as (Cu+)2S2 [18]. In addition, at this
pH, Fornasiero and Ralston, [14] also identied zinc and cupric suldes
as important species which were also responsible for the low zeta
potential. The sphalerite surface conditioned at pH 8.5 was dominated
by zinc and copper hydroxide/oxide and sulte/sulfate species. The
abundance of these species at pH 8.5 increases further with increasing
concentration of conditioning copper. These species are responsible
for making zeta potential values less negative above pH 5 and positive
above pH 6.5.
2.5. Copper concentration and activation duration
Fornasiero and Ralston [14] showed (under their reaction
conditions) that maximum collectorless otation recovery can be
obtained at copper concentrations of 2 10 6 moldm 3/g of
38 b x b 75 m sphalerite while Popov and Vucinic [15] found 3 10 5
to 2.5 10 5 moldm 3 Cu2+ concentration better for 104 b x b 208 m
sphalerite otation using xanthate. Typically copper concentrations
ranging from 1 10 4 to 1 10 6 M have been used to conduct
activation and otation studies of sphalerite [26,30,31,3744]. Section
3.3 provides typical plant concentrations used for copper activation.
The copper concentration needed to give maximum otation response
will differ depending on the activation conditions, origin of the mineral
sample and available surface area for activation. This ideal copper
concentration needs to be established through experimentation at
xed activating conditions. Fornasiero and Ralston [14] found that
increased Cu2+ concentration had the effect of continuously decreasing
the otation of sphalerite (in the absence of collector), with maximum
reduction seen at pH 8.510, probably due to the formation of Cu(OH)2
precipitates. The thickness of this Cu(OH)2 layer generally increases

101

with increasing copper concentration at mildly acidic conditions [18].


The effect of increasing copper concentration on otation is not evident
above pH 12 where Cu(OH)3 is the stable species (not Cu(OH)2 ppt)
and below pH 5 where only Cu2+ is the stable copper species [14].
High copper concentrations may also interact with xanthate
collectors in the pulp preventing adsorption onto the activated
sphalerite surfaces [15]. To minimise this side reaction, effectively
causing reagent loss, sequential addition (rst Cu2+, then collector
followed by lime) is now the preferred technique at Noranda Group
operations (Section 4.0) [11].
Prestidge et al. [18] found that an increase in activation time or
aging had the effect of reducing the Cu(OH)2 over-layer, evident from
the absence of high binding energy components and satellites from
the Cu(II) 2p XPS signal. They proposed that happens due to migration
of copper into the bulk sphalerite as surface Cu(OH)2 continues to
transform into the copper-substituted zinc sulde structure with
prolonged activation. Increasing the activation time therefore
increases the copper uptake with subsequent increase in the release
of Zn2+ [19] while increasing collector concentrations increases
collector adsorption onto the surface [4,5].
However, increased copper uptake may not be equally benecial
for surface xanthate adsorption Popov and Vucinic [26] conducted
oatability and infrared internal reection studies on a high iron
(13 wt.%) containing sphalerite (104 b x b 208 m) in weakly acidic
medium and investigated the effects of prolonged activation time and
copper concentration. For the short activation time (2 min) it was
found that Cu(I)xanthate was the dominant species on the surface
with a small concentration of Cu(II)xanthate. However, the amount
of Cu(I) xanthate decreased with increasing activation time, evident
from reduced peak intensity (at 1197 cm 1) in the IR spectra.
Prolonged activation time also resulted in a slow diffusion of Cu(I)
into the bulk sphalerite structure. Increased copper adsorption was
observed when the copper concentration was increased in solution
with long activation times, however Cu(I) xanthate formation was still
lower at than at short activation times. Interestingly, the otation
recovery of sphalerite was greater for samples activated for prolonged
periods and/or with increased activating copper concentration.
Lascelles et al. [30] demonstrated that an increased time delay
between Cu2+ addition and xanthate addition decreased the amount
of collector adsorbed on to the sphalerite surface at high pH (9.2) but
not at neutral or mildly acidic pH. At high pH it is known that Cu2+ can
also adsorb as Cu(OH)2 colloids [18]. On the basis of XPS measurements, Lascelles et al. [30] suggested that on delay between Cu2+ and
xanthate addition, colloidal Cu(OH)2 adsorbed onto the sphalerite
surface is lost to the solution resulting in decreased surface xanthate
adsorption. There was no (as was expected) decrease in the otation
response of sphalerite with reduced surface xanthate probably due to
collectorless otation, however the authors suggest early addition of
xanthate.
The presence of both copper and xanthate (and their quantity) may
also inuence interactions between sphalerite and gangue particles.
Duarte and Grano [39] conducted batch otation, zeta potential
measurements and rheological studies on silicate gangue minerals
(d85 of 1.0 m) and ultrane sphalerite (d85 of 7.9 m) to show that at
pH 9 silica can misreport to sphalerite concentrates through a
combination of entrainment and aggregation (with sphalerite particles) but not through true otation via surface hydrophobisation.
Under their reaction conditions, Duarte and Grano [39] found that
when no reagents were added in a mixed mineral system (silica and
sphalerite), silicate gangue minerals misreport to sphalerite concentrate primarily through entrainment. However, when 1800 g/t copper
sulfate is present with 1500 g/t isopropyl xanthate, silicate gangue
minerals misreport via a combination of entrainment and aggregation
as determined from batch otation and cryogenic SEM analysis. In
addition, rheological studies were used to show that particle
interactions increased upon copper sulfate and xanthate addition.

102

A.P. Chandra, A.R. Gerson / Advances in Colloid and Interface Science 145 (2009) 97110

They also found that the zeta potential of silica became less negative
upon addition of copper sulfate and xanthate. They theorised that
reagents modied the surface characteristics of silica causing the
surface charge to become less negative to values near the iep of silica,
where electrostatic repulsive forces are relatively low. This allowed
sphalerite and silica particles to interact and form aggregates that
report to the concentrates along with entrainment.
It has previously been shown that aggregation of sphalerite
particles can occur through occulation [36]. Mirnezami et al. [36],
on the basis of settling velocity, suspension analysis and optical
microscopy, showed the presence of aggregates in sphalerite particles
at pH 79 as a mechanism of sphalerite misreporting to lead and
copper concentrates in the processing of PbCuZn sulde ores. They
found that sphalerite releases sufcient zinc ions in solution (which
forms hydroxide in the pH range 79) and causes aggregation. A
occulating mechanism of aggregation was suggested involving
polymeric zero-charge Zn2+ species, [Zn(OH)2(H2O)2]0n.
2.6. The effect of sphalerite iron content on copper activation
Pure sphalerite (cubic ZnS) contains nearly 67 wt.% Zn and 33 wt.% S
[45]. However, natural ZnS normally contains iron (along with other
minor impurities) substituted for zinc atoms [46], amounts of which
depend on the temperature and chemistry of the crystallisation
environment [45,47]. The presence of iron decreases the band gap of
sphalerite, which is naturally an insulator, and affects its reactivity [48].
The iron content of sphalerite has been seen to inuence the activation
and subsequent otation behaviour of sphalerite during fundamental
studies; however contradictory results have been reported
[20,27,33,38,48,49]. Table 1 provides a summary of copper activation
and otation studies conducted over the last two decades with
sphalerite samples of various iron and lead (where this data is available
or lead is present) contents. This table also lists the various conditions
(where available) used for these studies. The pH, activator and collector
used are fairly consistent. However, other important factors such as
activation time and particle size are quite different.
Using synthetic sphalerite containing various iron contents (up to
40 wt.%) and 64Cu labelled CuSO4, Solecki et al. [27] showed that
adsorption of Cu2+decreased with increasing concentration of iron in
sphalerite. In another study, Szczypa et al. [49] further demonstrated,
with synthesised sphalerite, that increasing the iron content also
results in decreasing attachment of xanthate to copper-activated
sphalerite, primarily due to reduced copper on the sphalerite surface.
XPS studies by Buckley et al. [20] on two natural sphalerite samples
with different (high and low) iron contents also appear to support
these ndings. However, they acknowledged the gradient in lead

content of the two samples, where the sample with higher iron
content also had higher lead content, could also be exerting an effect.
Recent studies by Boulton et al. [38] also support the notion of iron
content of sphalerite reducing the rate of copper-activated sphalerite
otation. By carrying out otation studies (at pH 11) on two natural
sphalerite samples, with high iron (12.5 wt.%) and low iron (0.3 wt.%),
Boulton et al. [38] concluded that the presence of iron in the sphalerite
lattice reduces the exchange sites (zinc) for Cu2+, with this effect being
more pronounced for coarser particles presumably due to the lower
surface area to volume ratio. They however found that iron content
had no inuence on maximum recovery and otation rate constant at
low copper concentrations.
However, Gigowski et al. [33] and Harmer et al. [48] have reported
contrasting trends to those found by Solecki et al. [27], Szczypa et al.
[49], Buckley et al. [20] and Boulton et al. [38]. Using natural sphalerite
with varying iron content (up to 12 wt.%) they showed that copperactivated iron-rich sphalerite preferentially adsorbs xanthate. However, despite this no direct relationship between oatability, iron
content and copper concentration was found [33]. Recent studies by
Harmer et al. [48] demonstrated that as the iron content of sphalerite
increased the amount of Cu2+ adsorbing onto the sphalerite surface
also increased. Using a combination of electron microprobe analysis
(EPA), atomic force microscopy (AFM) and XPS on ve different
sphalerite samples with varying iron content (Table 1), they showed
that as the iron content of the sample increased the number of surface
defects and steps along with the size of surface oxidation products
also increased. The increased surface defect sites allow more Cu2+ to
be adsorbed compared to samples with low iron content (less defect
sites). In addition samples with higher iron content undergo a more
rapid oxidation than those with lower iron content, hence iron further
aids in Cu2+ adsorption. A decrease in the Fe 2p3/2 doublet intensity
was also noted compared that of the zinc, suggesting copper replaced
iron in the sphalerite preferentially over zinc.
2.7. Sphalerite surface oxidation
The sphalerite surface is characterised by steps and defects, size
and frequencies of which tends to increase with impurities such as
iron [48]. Therefore, sphalerite surface preparation for activation and
the degree of pre-oxidation of the surface may affect the available
surface area for the uptake of copper and xanthate, and are thus also
important factors to be considered. Harmer et al. [46] studied a high
iron sphalerite surface (110) using electron probe microanalysis
(EMPA), Rutherford backscattering (RBS), PIXE, XPS, and medium
energy ion scattering (MEIS) and found that vacuum fractured and air
fractured samples undergo varying degrees of relaxations and

Table 1
Copper activation and otation studies conducted over the past two decades and conclusions reached using sphalerite samples with different iron and lead contents
Sample (size)
Synthetic (not given)

Fe content

0%
5%
40%
Synthetic (not given)
0%
5%
40%
Natural (not given)
33a
115a
Natural (125200 m)
0.38%
12%
Synthetic: ZnSe and ZnS Natural: High (exact
Fe rich (b 10 m)
values not given)
Natural (range; 45 m)
0.3%
12.5%
Natural fresh (110)surface
0.02% to 14.79%
a

Pb content Activator
(time)

Collector used

pH

Conclusion

Reference

Not
present

64

CuSO4
(not given)

Increase in Fe causes decrease in


Cu adsorption

[27]

Not
present

CuSO4
(20 min)

Ethyl xanthate

6, 8, 10

Increase in Fe reduces xanthate


adsorption due to lower Cu adsorption

[49]

0.2a
24a
0.04%
0.20%
(Not
given)
0.07%
0.24%
(Not
given)

CuSO4
(1 h)
CuSO4
(not given)
CuSO4
(1 min)
CuSO4
(2 min)
CuSO4
(1 h)

9.2

[20]

Sodium isopropyl
xanthate
Sodium isopropyl
xanthate
Sodium isopropyl
xanthate

Cu uptake by high Fe sample was less


than for the low Fe sample
Fe rich sphalerite adsorbs more Cu and
preferentially binds to xanthate
Fe inhibits initial Cu activation but catalyses
covellite formation upon xanthate addition
Fe is detrimental to copper activation which
reduces collector adsorption
Fe enhances Cu adsorption

Values reported as atomic ratios ( 103) of metallic impurity elements relative to zinc.

10 and
12
11
5

[33]
[17]
[38]
[48]

A.P. Chandra, A.R. Gerson / Advances in Colloid and Interface Science 145 (2009) 97110

reconstructions resulting in possible SS (dimer) type bonding.


However these surfaces are likely to be considerably different to
those exposed to an aqueous environment.
Leaching studies conducted by Weisener et al. [50] have shown
iron-rich sphalerite to be more oxidised and to leach more rapidly
than low iron containing sphalerite, suggesting greater reactivity of
iron-rich sphalerite surfaces. Recent AFM images have shown the
occurrence of larger and more frequent oxidation features on air
exposed high iron sphalerites (compared to low iron sphalerite),
conrming higher reactivity of high iron sphalerite [48].
Solecki et al. [27] established that surface oxidation prior to
activation had a greater effect on copper adsorption onto the low iron
sphalerite in comparison to the high iron sphalerite. Szczypa et al. [49]
demonstrated that the copper/xanthate uptake was much greater
onto unoxidised than oxidised sphalerite surfaces for both high and
low iron sphalerite and that the effect of iron content was more
pronounced for unoxidised as compared to oxidised surfaces.
Gigowski et al. [33] however found that copper activation of sphalerite
surfaces was more inuenced by iron content than by the degree of
oxidation. These results suggest that while both increased oxidation
and increased iron content are detrimental to otation it is the latter
which has the more profound effect.
Using XAS Pattrick et al. [17] showed that activated natural (Zn,Fe)S
was more oxidised than ZnS, and that surface iron in (Zn,Fe)S existed
as FeO. Similar FeO structures were observed by Buckley et al. [20] on
freshly fractured sphalerite samples, which were believed to have
existed prior to the fracture. According to Pattrick et al. [17], the
existence of FeO had the effect of accelerating the formation of CuO
and covellite (CuS), which is proposed by some authors to be among
the major products of activation and xanthate addition. Additionally,
the EXAFS Fe K edge data revealed that iron also reacts with the sulfur
of xanthates forming FeS, thus adding further evidence to the notion
of preferential xanthate adsorption on high iron sphalerite. These
ndings also provide a possible explanation for the observation made
by Gigowski et al. [33].
2.8. Lead and iron sphalerite activation
Sphalerite containing a high iron concentration also generally has
higher lead content. Addition of, or the presence of, lead ions (Pb2+) in
the otation slurry can directly activate sphalerite and promote
otation [5153]. Pb2+ can be present due to dissolution from galena
and also through recycling of plant water, which can lead to
inadvertent lead activation and misreporting of sphalerite to copper
and lead concentrates [11,54]. The presence of greater lead content in
high iron sphalerite (compared to low iron sphalerite) is likely to
contribute towards higher otation response seen in studies with high
iron sphalerite (Section 2.6).
Pb2+ on the sphalerite surface (either present naturally as part of
sphalerite lattice or adsorbed from the solution) promotes otation by
reacting with xanthate forming stable xanthate and/or dixanthogen
[55]. Basilio et al. [55] using kinetic studies and spectroscopic analysis
(XPS, XRF and FTIR) suggested that lead activation of sphalerite occurred
through an exchange mechanism involving exchange of Pb2+ with Zn2+
(similar to copper activation) on the sphalerite surface. This exchange
mechanism is however not supported by other studies as accommodating lead in place of zinc sites would require considerable relaxations of
local structure as lead is much larger than zinc (or copper) [53,56].
Computer (atomistic) simulations by Pattrick et al. [56] showed
that lead is incompatible in zinc sites in a sphalerite lattice (both
surface and bulk) and the exchange process is not energetically
favourable. Further uorescence REFLEXAFS study by Pattrick et al.
[56] favoured a mechanism involving the formation of Pb-oxide
species which become a point of attachment of xanthate sulfur to the
lead. A high pH (9.2) normally favours more lead adsorption on
sphalerite surface and it may occur through direct adsorption

103

(without exchange) of Pb2+ followed by development and precipitation of lead hydroxy species such as PbOH+ and Pb(OH)2 [51,53]. In
addition, carbonate species such as Pb3(CO3)2(OH)2 precipitates can
also occur at high pHs and under prolonged exposure to air, which is
typical in processing circuits. This is evident from solution speciation
calculations of O'Dea et al. [57]. The Pb(OH)2 may then react with
xanthate to form lead xanthate through an ion exchange mechanism
[51]. Rashchi et al. [52] however found through micro-otation tests
that lead activation of sphalerite was signicant below pH 7 and
progressively decreased to zero at pH 11. Based on their results and
those from the literature, Rashchi et al. [52] suggested a mechanism
where lead exchanges with zinc in the sphalerite lattice and reacts
with xanthate to form PbX only in the acidic pH range. From pH 710
lead forms a ZnOPb+ species on the surface through adsorption of
Pb(OH)+ but also forms the same PbX species with xanthate. Beyond
pH 10 the Pb(OH)2 precipitates dominates which can render sphalerite
surface hydrophilic thus depressing otation.
The mechanism of lead activation of sphalerite remains poorly
understood and further surface studies involving both spectroscopic
(XPS, ToF-SIMS, NEXAFS, EXAFS) and electrochemical kinetic studies
are needed for clarication. Synchrotron based measurements
(spatially resolved) may be better suited as the surface lead
concentrations are relatively low and may involve localised heterogeneous adsorption.
Additions of Fe2+ in the presence of oxygen can also activate the
sphalerite surfaces (through adsorption of Fe2+ as Fe(OH)+ followed by
anodic oxidation to Fe(OH)2+) and aid sphalerite oatation by forming
a ferric hydroxy complex with collector molecules at moderately
alkaline pH [34,58]. Such results with ferrous activation via aqueous
addition may provide insight regarding the possible inuence of iron
in the sphalerite structure (bulk and surface) with evidence of FeO [17]
type structures on the sphalerite surface and also the possible
inuence of the presence of iron containing minerals such as pyrite,
which can contribute to the pulp iron concentration through
solubilisation. The use of steel grinding media however contributes
to a larger portion of pulp iron content through oxidation of steel
(corrosion) during grinding (Section 4.0).
However, iron or lead within the sphalerite lattice do not appear to
signicantly inuence collector adsorption (through solubilisation
during conditioning) under laboratory conditions. Tong et al. [59]
conducted unactivated (without copper) otation tests (2 min collection time) on high iron marmatite with varying concentrations of
butyl xanthate. The recoveries obtained for all xanthate concentrations used were extremely low with percent recoveries less than 5%.
Similar results were also found by Boulton et al. [38] for ZnS and (Zn,
Fe)S using the collector SIPX. Zhang et al. [34] found that activation
effect of Fe2+ in the pulp solution was more pronounced between 1
and 2 ppm and that their effect continuously decreased above 2 ppm.
They also did not nd any activation effect by Fe3+. Under laboratory
conditions of activation, iron or lead within the sphalerite lattice may
not be solubilising sufciently to produce similar effects to those seen
in studies with aqueous additions of these ions into the mineral slurry.
Within a otation circuit these ions (lead and iron) possibly occur
as a result of dissolution from minerals that predominantly contain
these elements. Table 2 shows rest potential values of some common
Table 2
Rest potential of some common sulde minerals
Sulde mineral

Rest potential (SHE) V

Pyrite
Chalcopyrite
Sphalerite
Covellite
Bornite
Galena

0.66
0.56
0.46
0.45
0.42
0.40

Obtained from [60].

104

A.P. Chandra, A.R. Gerson / Advances in Colloid and Interface Science 145 (2009) 97110

sulde minerals [60]. In a mixed mineral system (as in plants) galvanic


interactions occur, with minerals of higher rest potential being
cathodic and those with lower rest potential being anodic. Anodic
mineral are prone to oxidation unlike cathodic minerals. Galena, with
the lowest rest potential, will undergo oxidative dissolution in a mixed
mineral system and may be a signicant source of Pb2+ in otation
circuits. Pyrite will be cathodic and will not contribute as much iron to
the system. The source of iron in mineral circuits may be from other
sources such steel grinding media (Section 4.0).
Zielinski et al. [61], using SEM-EDS analysis, found evidence of
preferential misreporting only of low-iron sphalerite into lead
concentrates within three Cominco lead otation circuits, most likely
through activated otation with collector adsorption. Its was
envisaged that the low-iron particles (both coarse and ne size
fractions) were activated possibly by lead solubilised from the ore
within the slurry and that preferential activation of only low iron
sphalerite particles could have been due to greater oxidation of high
iron sphalerite particles due to their higher reactivity. Such studies
highlight the complexity of the activation and otation process used
for mineral separation in actual plant environment which may be
missed under controlled laboratory conditions.
3. Copper activation of pyrite
Unlike sphalerite, pyrite responds well to thiol collector molecules
in the absence of activation [4,6], however pyrite can be activated
when copper is present in the slurry. In either case, inadvertent pyrite
otation can result, either during otation of copper bearing
minerals/ores due to partial dissolution [62] or during activated
otation of sphalerite especially where the concentration of sphalerite in the slurry is low [6]. Secondary copper bearing ores such as
chalcocite, covellite and bornite pose greater risk of inadvertent pyrite
activation and otation compared to chalcopyrite, as evident from
rest potential values in Table 2 [60,62,63]. It is therefore important to
understand the controlling mechanisms of pyrite activation and the
otation of pyrite into copper or zinc concentrates. In addition to Cu2+,
pyrite can also be activated by Pb2+, Fe2+ and Ca2+ which can lead to
inadvertent pyrite otation causing it to misreport to copper, zinc or
lead concentrates [6,64,65]. As for sphalerite, pyrite activation and
otation is also inuenced by pH, copper concentration and activation
time.
3.1. Unactivated otation
Pyrite has a natural otability and a small percentage of it can
usually be recovered without any use of activators or collectors
[40,65]. Much higher recoveries can be achieved with collectors. It has
been shown by Leppinen [4] that adsorption of xanthate onto the

pyrite surface without activation reaches a maximum at pH 5 and a


minimum at pH 7. Flotation results showed that as much as 80 to 90%
of pyrite can be recovered through otation (without activation) at pH
4 to 5, while recovery is much lower at other pH values [4,6]. Surface
analysis by Leppinen [4], of unactivated pyrite samples conditioned
with ethyl xanthate conrmed the occurrence of ironxanthate with
diethyl dixanthogen at monolayer coverage on the pyrite surface.
A clear distinction as to whether the ironxanthate is in the Fe(II)
or Fe(III) form was not made by Leppinen [4] however some Fe(III)
character was evident. Studies by Valdivieso et al. [66] have shown
that adsorption of xanthate onto unactivated pyrite surface increases
aqueous Fe2+. The surface oxidation of xanthate to dixanthogen results
in a corresponding reduction of the surface Fe(III) hydroxide with
conversion to Fe2+. This mechanism proposed by Valdivieso et al. [66]
is shown in Fig. 2. According to this mechanism as the hydrophobic
dixanthogen develops on the pyrite surface there is a subsequent
reduction in hydrophilic surface hydroxide.
3.2. Mechanisms of Cu(II) and Cu(OH)2 activation
There are important structural and electronic differences between
pyrite and sphalerite and as such it is no surprise that the copper
activation of pyrite follows a different mechanism from that of
sphalerite. It was shown by Weisener and Gerson [67] that copper
uptake during pyrite activation does not results in a related 1:1 (or
any) iron release from pyrite, thus ruling out an ion exchange
activation mechanism. Pyrite activation follows a single fast step
involving Cu(II) adsorption onto the reactive sulfur sites only on the
surface with no migration into the bulk pyrite [67]. During adsorption
Cu(II) is reduced to Cu(I) with subsequent oxidation of surface sulde.
Boulton et al. [37], conrmed that copper adsorbs only to the surface
and does not migrate to the bulk, as is case for sphalerite.
Using Fourier transform infrared spectroscopy employing attenuated total reection (FTIR-ATR), Leppinen [4] suggested that the pyrite
surface upon activation is fully covered with coppersulde like
products which resemble Cu2S more than CuS. However, subsequent
EXAFS data analysis has indicated that copper on the activated pyrite
surface has a distorted trigonal planar position between three sulfur
atoms [67] with an average CuS bond length of 2.27 0.02 .
Weisener and Gerson [31,67] studied pyrite activation using
EXAFS, XPS, angle resolved XPS and time of ight secondary ion
mass spectrometry (ToF-SIMS) and found Cu(I) to be present for all pH
and all copper concentrations studied, with Cu(II) occurring (as
hydroxide) and overlaying the Cu(I) activated surface only at alkaline
pH. Moreover, no further copper uptake was observed after 5 min of
activation. At alkaline pH, along with the distorted trigonal planar
copper moiety, Cu(OH)2 precipitates were also found with a CuO
bond length of 2.00 .

Fig. 2. Adsorption and dixanthogen formation on unactivated pyrite surface. Redrawn from [66].

A.P. Chandra, A.R. Gerson / Advances in Colloid and Interface Science 145 (2009) 97110

Zhang et al. [6] suggested through zeta potential and FTIR (ATR)
measurements that copper is chemisorbed to the surface. It was
however shown by Hicyilmaz et al. [68] that the interaction between
copper and pyrite is solely an electrochemical process while the
interaction between activated copper and collector, and the pyrite
surface and collector is primarily chemical in nature. Results from
Pecina et al. [65] have also conrmed the chemical interaction of
collectors with activated and unactivated pyrite surfaces.
Leppinen [4], through IR internal reection surface analysis of
activated pyrite surfaces conditioned with ethyl xanthate collector,
showed Cu(I)ethyl xanthate to be the dominant activation product
with monolayer coverage. Shen et al. [69] also reported cuprous
xanthate as the only xanthate product formed under their reactions
conditions. Leppinen [4] found copper concentration and pH
dependency of xanthate adsorption and subsequent otation of
pyrite. They showed at pH 7 that when equal amounts of copper
and xanthate are used, copper xanthate is the only product formed,
however when either relatively higher or lower concentrations of
copper is used, signicant amounts of dixanthogen also forms. Some
Cu(II)xanthate like compounds were also observed at higher copper
concentrations. Oxidation of xanthate to dixanthogen has also been
reported by [6].
3.3. Effectiveness of activation
The amount of xanthate adsorbed on activated pyrite surfaces was
seen to increase from a minimum at pH 45 to a maximum at pH 8 [4].
The adsorption decreased drastically above pH 8. This can also be
observed from the data of Zhang et al. [6] and Dichmann and Finch
[70] where percentage recovery of activated and unactivated pyrite,
with and without xanthate, undergoes a similar decrease after pH 89.
They however obtained a signicantly higher recovery than Leppinen
[4] at pH 4.
Generally, activation by copper results in signicant increases in
recovery only within the pH range of 610 [4,6,70]. At mildly acidic pH
a higher than expected recovery of unactivated pyrite is observed
owing, most probably, to the emergence of sulfur-rich products (due
to the dissolution of iron). Hence, in these studies, the maximum
otation, using xanthate collector, of unactivated pyrite at pH 4 is seen
to exceed otation of activated pyrite at pH 8. This may have very
limited practical implications as most plants don't operate at a low pH
of 4. It is generally seen that hydrophobicity of pyrite surfaces
increases at low pH (compared to higher pH under similar conditions)
with either activation or collector addition or both [68], thus resulting
in higher recovery. Flotation of both activated and unactivated pyrite
generally follows rst-order kinetics [63].
He et al. [40] have shown pulp oxidation potential (Eh) to be an
important factor in determining recoveries and speciation on pyrite
surfaces, with maximum recoveries obtainable at the conditioning
oxidation potential of 35 mV (SHE) at pH 9. EDTA extraction and
surface studies (XPS) revealed that Eh inuences the production of
hydrophilic (iron oxide/hydroxide) and hydrophobic (Cu(I)S) species,
and also promotes formation of Cu(I)xanthate species on pyrite
surface The presence and relative abundance of such surface species
has a corresponding effect on pyrite recovery.
Typically copper and xanthate concentrations from 1 10 4 to
1 10 6 M are used to conduct fundamental lab-based studies
[14,26,30,31,3744]. A few studies have also reported using 100
1500 g/t collector [3739,4143] and 2503000 g/t copper containing
activator (such as nitrate or sulphate) [3739,42,70]. Table 3 gives
names of some collectors and their typical addition rate along with the
dose of copper sulfate used in otation of sulde ores in processing
plants. A range of collector and copper concentrations is used
depending on the exact mineralogy of the ore being processed and
the target minerals. It is however difcult to compare the values given
in literature, which are normally expressed in molar concentrations,

105

Table 3
Typical amounts of collector and activator used in otation and separation of sulde
ores at processing plants
Reagent

Name

Amount used (g/t)

Collector

Xanthate
Dithiophosphate
Thionocarbamate
Xanthogen formate
Xanthic ester
Mercaptobenzothiazole
Thiocarbanilide
Copper sulfate

5350
10250
1015
225
225
25250
2575
1002500

Activator
Data obtained from [9].

and plant values. Most of these studies do not state the conditioning
volume, hence it becomes difcult to convert molar concentrations
into g/t. From the studies that do express reagent concentration in g/t,
the collector concentrations used appear to be an order of magnitude
higher than normal plant practice. Studies beyond the normal
practical range are however needed to provide new perspectives for
better separation. With exception of a few studies, most references
quoted have conducted tests on single mineral system while
maintaining concentrations similar to that used in plant. Needless to
say that processing plants have mixed mineral systems with galvanic
processes which does not occur in single mineral system.
4. Mixed pyrite and sphalerite otation
Separation of pyrite and sphalerite through activated otation is
normally carried out at high pH [3,71] although some plants such as
Teck Cominco also does this at lower pH [48]. Pyrite has a higher rest
potential than sphalerite and due to this pyrite surfaces become
coated with OH products (resulting from reduction of O2) due to
galvanic coupling, making pyrite surface less hydrophobic, thus
increasing sphalerite selectivity [72]. This galvanic coupling can
however be suppressed by saturating the ore slurry with N2 gas and
using N2 gas for otation instead of air. This causes the pulp potential
to be reduced due to reduced oxygen activity [72]. The use of N2 gas to
reduce galvanic interactions can be used in reverse otation of pyrite
from sphalerite [11]. It is also used effectively in the N2TEC otation
technology for otation of auriferous pyrite by the Newmont Mining
Corporation at their Lone Tree Plant in Nevada [64]. This process was
developed initially to improve recovery of gold from low grade sulde
ores, the N2TEC process involves processing of ores (from grinding to
otation) in an inert N2 gas environment where operating potential
ranges from 100 mV to 300 mV vs. SHE (0.1 to 0.5 V vs. Ag/AgCl) and
uses potassium ethyl xanthate [64,73].
Zhang et al. [6] conducted micro-otation studies of activated
pyrite in the absence and presence of sphalerite. Activated pyrite
otation is observed to be depressed signicantly in the presence of
sphalerite at all pH values with recovery continuously decreasing to
nearly 2% at pH 11. Pyrite recovery at pH 4 is seen to be much higher
than recoveries at all other pH even when conducted in the presence
of sphalerite. Flotation without copper activation generally has no
effect on pyrite recovery as sphalerite does not combine with xanthate
without copper [6].
Using micro-otation, laboratory batch, and in-plant minicell
experiments, Dichmann and Finch [70] observed similar behaviour
during combined otation of activated pyrite and sphalerite. From
these studies it appeared that during mixed activation/otation
sphalerite preferentially consumed copper and xanthate while pyrite
becomes depressed when in the presence of sphalerite. According to
Dichmann and Finch [70], addition of copper increases galvanic
coupling between sphalerite and pyrite particles, and this favours
xanthate adsorption on sphalerite while pyrite becomes coated with
hydrophilic hydroxide ions. The effectiveness of this electrochemical

106

A.P. Chandra, A.R. Gerson / Advances in Colloid and Interface Science 145 (2009) 97110

reaction in the separation of sphalerite from pyrite is still not fully


understood and remains an area of continued research [74]. However,
according to Finch et al. [11], the notion of pyrite depression upon
copper addition, by increased galvanic coupling with sphalerite, has
successfully been tested at two plant operations in Canada, Agnico
Eagle's Laronde and Noranda's Matagami-Bell Allard. In these tests
copper and lime were added sequentially to initiate selective copper
activation of sphalerite rather than using traditional method of
simultaneous addition of copper and lime or lime addition prior to
copper addition. This has now become part of Noranda operations
after a signicant reduction of iron from zinc concentrates was
achieved.
Recent surface analytical studies, XPS and ToF-SIMS, with mixed
pyrite and sphalerite otation have shown evidence of hydrophilic
species such as ferric hydroxide/sulfate obscuring the pyrite surface
[37]. This iron hydroxide layer appeared to have inhibited copper and
collector adsorption onto the pyrite, thus resulting in a decrease in
pyrite recovery. Hydrophobic species like cuprous sulde and
collector were on-the-other-hand dominate on the sphalerite surface
[37]. He et al. [40] also observed the inhibiting iron oxidation species
on pyrite, leading to a decreased recovery. Boulton et al. [37] further
established that separation of these minerals depends on the
concentration of activating copper and concentration of collector
used. Increasing copper concentration was found to increase sphalerite recovery while increasing collector concentration increased pyrite
recovery. The concentrations of copper and xanthate should be aimed
to be kept at a level which gives maximum otation response from
sphalerite while still retarding pyrite otation into zinc concentrates.
Particle size may also have an effect on the otation response of the
sphalerite when mixed with pyrite. Boulton et al. [37] further found
that ner sphalerite particles report to the tails due to the lower
probability of such particles combining with gas particles during
otation.
Further selectivity can be achieved by carefully choosing and using
different collectors and otation conditions. Shen et al. [71] found
ethoxycarbonyl thionocarbamate and thiourea collectors to be more
selective for separation of copper-activated sphalerite from pyrite
than xanthate in the neutral to alkaline pH range. Moreover, they also
found that the use of O2 gas, for conditioning, improved selectivity
towards sphalerite separation from pyrite especially at short mineral
conditioning times of less than 20 min. Galvanic interactions between
mineral particles are reduced to some degree at high pH as the pulp
potential is lowered [72]. Pulp potentials and hence galvanic reactions
can be restored by conditioning with air or O2, thus increasing
sphalerite selectivity. Excessive oxygenation should however be
avoided as otability by thiol collectors can be affected [72]. Grinding
with the use of mild steel also reduces pulp potential as corrosion of
steel consumes O2. The corrosion of the grinding media is accelerated
by presence of galvanic interactions between the steel media (lowest
rest potential) and mineral particles, with cathodic reduction of O2
occurring on mineral surfaces [9,42]. This not only leaves mineral
surfaces rich in hydrophilic hydroxyl products which affect selectivity
but is also a signicant source of iron contamination [11]. Grinding
with inert media (sand, slag, ore or ceramic) is preferred over steel
[11]. Modications were made in the otation circuit at Mt. Isa Mine
Ltd., Australia, to obtain better zinc concentrate grades through inert
regrinding of leadzinc ore at key streams to liberate more sphalerite
through ner grinding [9,75].
Separation efciencies can also be increased by the use of sulfoxy
species such as sodium sulte [69], SO2 or sodium metabisulte [72].
Sulfoxy species acts as depressants and they depress minerals like
pyrite by preventing collectors from adsorbing onto the mineral
surface. Rao [72] suggests that since no sulfuroxygen products are
observed on pyrite surfaces, depression by sulfoxy species tends to
follow an electrochemical mechanism. Using a modied Smith and
Partridge micro-otation column, Shen et al. [69] obtained maximum

separation of copper-activated sphalerite from pyrite in the presence


of sodium sulte at pH 8.5 and separations were much better when O2
was used as the conditioning gas and when the time interval between
additions of sulte and xanthate were relatively short. Fig. 3 shows
comparative results from this study with varying intervals between
xanthate and sodium sulte addition (20 s and 10 min) and the use of
O2 or N2 as conditioning gas. Sodium sulte and O2 conditioning is also
seen to increase separation efciencies through pyrite depression
rather than through an increase in sphalerite otation. A 70 mV
decrease in solution Eh was also observed upon sodium sulte
addition. Based on these results and combined UVvisible, IR and XPS
studies, Shen et al. [69] proposed an electrochemical mechanism of
pyrite depression with the formation of copper hydroxide on copperactivated pyrite surfaces through consumption of O2 which causes the
decrease seen in solution Eh. The use of sodium sulte and O2
conditioning helps depress copper-activated pyrite surfaces through
the following mechanism:
2

1
SO2
3aq 2O2aq YSO4aq 2e

12

H2 Oaq 2e 12O2aq Y2OHaq

13

Cu2
aq 2OHaq CuOH2surface

14

In addition to this mechanism Shen et al. [69] propose that galvanic


interactions occur between the cuprous sulde layer on the pyrite
surface and the pyrite mineral surface. The cuprous sulde layer being
less cathodic is oxidised to produce cupric ions while O2 is reduced at
the pyrite surface producing hydroxide ions. A similar process also
occurs on the sphalerite surface however it occurs predominantly on
pyrite as it is the more cathodic of the two minerals. Hence, sodium
sulte with O2 induces more hydroxide formation on pyrite surfaces
than on sphalerite.
A range of sulfoxy species from S2 to SO2
3 can be used as
depressants, and the use of metabisulte in addition to controlling
electrochemical conditions (Eh and pH), has resulted in signicant
pyrite depression in-plant trials at Heath Steele for oating galena
away from pyrite [11]. Sulfoxy species sodium sulde is also used in
the suldisation of oxidised sulde mineral surfaces to restore
weathered surfaces and to retard metal ion dissolution (such as
lead) that can inadvertently activate pyrite [11,72]. The sulfoxy species
also sequesters metal ions already released into the solution. This
reduces the concentration of xanthate needed to oat oxidised
minerals as side precipitation of metalxanthate is reduced.
Polyphosphates has also been found to prevent accidental
activation of unwanted minerals by metal ions such as lead present
in the mineral slurry [76]. Micro-otation tests by Rashchi and Finch
[76] showed that sodium polyphosphates effectively deactivated

Fig. 3. Sphalerite and pyrite recoveries obtained at pH 8.5 with [Cu2+] and [KEX] =
2 10 5 moldm 3 and [Na2SO3] = 2 10 4 moldm 3. Adopted from [69].

A.P. Chandra, A.R. Gerson / Advances in Colloid and Interface Science 145 (2009) 97110

sphalerite in a sphaleritelead otation at near neutral conditions.


Based on SEM and XPS studies a cleaning mechanism of polyphosphate action was suggested which involved removal of activating
lead ions from the sphalerite surface by formation of a soluble
complex. The use of phosphate as a lead selective deactivator greatly
improved copper concentrates at NVI Mining's Myra Falls operation by
reducing the amount of zinc and lead in the concentrates [54].
Sodium cyanide can also effectively depress pyrite by inhibiting
xanthate adsorption and its subsequent oxidation [72,77]. This
happens through the formation of an insoluble ironcyanide complex
through an electrochemical mechanism [44,78]. Cyanide also displaces xanthate already adsorbed on to pyrite surfaces by an exchange
mechanism [44]. Organic depressants such as dextrin, starch, diethylene-triamine and tri-ethylene-triamine (DETA/TETA) are being
considered as a more environmentally friendly option to cyanide as
depressant [11,79,80]. In addition cyanide has been found to be
responsible for inadvertent activation of sphalerite (by Cu) during
otation of chalcopyrite, which had cyanide added to depress pyrite
[81]. Mining plant in Quebec, Canada, suffers from this inadvertent
activation and otation of sphalerite into chalcopyrite concentrates
when using cyanide as a depressant. Studies by Rao et al. [81]
concluded that Cu is leached out of chalcopyrite by the cyanide and
under sufciently anodic pulp potentials inadvertent activation
of sphalerite takes place. However when the pulp potential is cathodic
(b 155 mV SHE) this inadvertent activation is suppressed as the
solution Cu exists in Cu+ oxidation state and is not able to exchange
with divalent zinc from sphalerite.
Another important factor that can impact the selectivity for
sphalerite and pyrite otation is the quality of water used in-plant
otation. The primary water source for mine usage can be relatively
impure. Treated sewerage and hard water are regularly sourced which
can have varying organic and inorganic content [82]. In addition the
process water in-plant otation is increasingly recycled from tailings
dams, thickener overows, dewatered and lter products, which can
lead to a build up of residual reagents and reaction products (such as
xanthates and dixanthogen), mineral dissolution and oxidation
products (metal ions such as Fe2+, Cu2+, Pb2+, Na+, K+, Ca2+ and Mg2+),
colloidal precipitates, residual suldes, increased microbiological
activity and highly variable pH and Eh [8284]. Such degradation of
water quality tends to affect otation and selectivity of both sulde
and non-sulde mineral systems and increases reagent consumption
[8388]. Depending on water quality otation and selectivity of
sphalerite and pyrite (and other sulde systems) can be affected by
non-selective adsorption of xanthate and dixanthogen, depression
from accumulated suldes/cyanides and inadvertent activation from
metallic ions such like Fe2+, Cu2+, Pb2+ and their colloidal precipitates.
Water quality in otation circuits can be improved by employing
waste water treatment techniques such as dissolved air otation
(DAF), organic removal through biological oxidation, occulation and
aggregation, precipitation, ion exchange, UV exposure and reverse
osmosis [82,83,89]. Water quality is not a problem in lab-based
fundamental studies as distilled, de-ionised or de-mineralised water
are frequently used, which may not approximate in-plant (electrochemical) conditions. As such outcomes from fundamental studies
may not equate to similar outcomes in an actual processing
environment. Fundamental studies however provide the basic
chemical and kinetic information that is invaluable and plant trials
where possible can add further value.
5. Summary
A wide range of studies on sphalerite and pyrite otation has been
conducted over several decades, which range from the investigation of
basic mineralogy to elucidating complex surface structures, with the
overriding aim of providing fundamental knowledge to enable
increased economic benets. The information from such studies

107

needs to be consolidated as a feed back system to realise the extent of


progress made and the milestones that are yet to be achieved. As
evident from the literature reviewed there are three general areas
which require further attention. The rst is to recognise and be able to
predict the optimum conditions needed to activate, oat and
effectively separate pyrite and sphalerite. The second area is to
determine the structures formed after activation and collector
addition and the mechanism leading to these species. And nally an
effective way to reconcile these lab-based studies with plant practices
is required to ensure maximised process throughput, efciency and
economy and minimised environmental impact.
Both sphalerite and pyrite can be activated by ions such as Cu2+, Fe2+
and Pb2+ in solution; however it is only copper that is used
commercially while the other ions are considered to be contaminants
which can cause inadvertent activation and loss of selectivity. Copper
activation of sphalerite proceeds with a 1:1 exchange of Cu2+ with Zn2+
from the rst couple of atomic layers of the sphalerite surface after
which the related sulde is oxidised to a Cu(I)S species. This step is
relatively rapid (completed within rst 1520 min) with the rate most
likely controlled by solution mass transfer, thus requiring sufcient
stirring during conditioning to maximise the rate of activation.
After the initial step, there is a slow steady state counter diffusion of
copper (most likely in the Cu(I) oxidation state) into the bulk structure
and Zn2+ out of the bulk structure. The surface copper sulde species
formed is hydrophobic and can induce collectorless otation especially
under low pH conditions. Pyrite activation on-the-other-hand follows
a single fast step involving copper adsorption to the surface without
any exchange with iron in the pyrite lattice. Copper adsorbed onto the
pyrite surface does not migrate into the bulk pyrite lattice.
Unlike sphalerite pyrite responds to thiol collectors and tends to
oat well even without copper activation. This behaviour is due to the
activating nature of Fe2+ which is naturally present on pyrite surfaces,
where ironxanthate structures form along with dixanthogen. This
largely happens under low pH conditions as the surface Fe2+ sites
form iron hydroxide species at higher pH which retards collector
adsorption.
Cu(I)xanthate (or any other collector) is the dominant species
formed when sphalerite and pyrite are individually activated and
conditioned with xanthate collector. Under conditions (such as long
activation time or increased delay between copper activator and
collector addition) which allow copper to migrate into the sphalerite
lattice, the amounts of Cu(I)xanthate formed on the surface will be
relatively low, which depending upon other slurry conditions may or
may not affect otation. Xanthate on the mineral (sphalerite and
pyrite) surface or in the solution may also dimerize to form
hydrophobic dixanthogen especially at low pH and relatively low
copper concentrations. Dixanthogen formed in solution tends to
adsorb non-selectively and can cause unwanted otation of minerals
especially gangue. Xanthate may also react with Cu2+ in the
conditioning solution causing precipitation of Cuxanthate and
again non-specic adsorption. This limits the amount of xanthate
available for oating the desired mineral thus increasing reagent
consumption. Sequential addition of copper and xanthate is consequently preferred over simultaneous additions of these reagents.
At high pH conditions colloidal Cu(OH)2 will also precipitate onto
mineral surfaces and a thick obscuring layer of Cu(OH)2 may form if
the conditioning copper concentration is relatively high. The presence
of Cu(OH)2 on the surface can cause loss of selectivity as Cu(OH)2
precipitates non-selectively, hence has the potential to cause otation
of unwanted mineral or gangue, through reaction with xanthate. The
Cu(OH)2 layer also obscures the Cu(I)sulde layer thus further
reducing surface hydrophobicity. Cu(OH)2 precipitated on sphalerite
surface may undergo ion exchange with surface Zn2+ followed by
reduction to Cu(I). With time (or activation time) this Cu(OH)2 will
continue substituting with Zn and will slowly migrate into the bulk.
This exchange does not take place between iron on the pyrite surface

108

A.P. Chandra, A.R. Gerson / Advances in Colloid and Interface Science 145 (2009) 97110

and adsorbed Cu(OH)2. Some Cu(OH)2 precipitates will also be lost


back into the solution especially if the solution copper concentration
falls below Cu(OH)2 saturation. The effect of Cu(OH)2 will however
depend on conditioning copper and xanthate concentrations, conditioning time and pH. In the case of pyrite higher pH tends to
promote formation of iron hydroxide and copper hydroxide species
which not only block collector adsorption sites but also makes the
surface more hydrophilic thus reducing the otation response. Higher
pH is ideal for depressing pyrite while lower pH can be used for
oating auriferous pyrite.
While copper concentration, activation time and pH have clearly
been shown to affect sphalerite activation and otation, the iron
content of sphalerite has been a point of controversy, conned
however only to fundamental studies. Iron content of sphalerite is
currently not used as a basis for sphalerite processing in plants. It is
likely that the presence of iron in the sphalerite lattice reduces its
band gap and increases its reactivity by aiding in electron transfer
reactions. However, the presence of a high lattice iron concentration
has been shown to give rise to increased adsorbed copper and
subsequently higher xanthate adsorption but also lower adsorbed
copper and xanthate adsorption. The reason for this discrepancy lies
entirely on the higher reactivity imparted to sphalerite by presence of
high iron. Sphalerite containing high iron oxidises much faster than
low iron sphalerite and at high pH is likely to be covered by
hydroxides possibly at iron sites which may be located near steps or
defects. Surface pre-oxidation and pH is therefore critical. The effect of
high iron content on sphalerite copper adsorption, in the absence of
surface oxidation, can only be clearly demonstrated at low pH using
samples cleaved (or ground) in situ or in an inert environment. Low
pH ensures that there are no obscuring hydroxides on the surface
while in situ cleaving or cleaving under an inert environment ensures
no atmospheric oxidation of the surface.
Apart from enhancing electrochemical reactions, iron may aid in
copper adsorption by preferentially (over zinc) exchanging with
copper and may also combine directly with sulfur in xanthate. Under
conditions of high pH or conditions which do not limit atmospheric
surface oxidation, surface reactive sites become obscured by oxidation
products including hydroxide species. Under such condition low iron
sphalerite adsorbs more copper or xanthate due entirely to its less
reactive nature. This can cause sphalerite with low iron content to
misreport to copper or lead concentrates. While the notion of
sphalerite reactivity affecting its activation by copper is plausible, it
does not however rule out inuence from other factors such as
activating copper concentration or presence of contaminants such as
lead. The concentration of lead in sphalerite tends to increase with
increasing iron concentration however this factor is frequently not
taken into account. While the presence of contaminants such as lead
in process water is usually not an issue for lab-based single mineral
studies, long activation times may however be sufcient to solubilise
lead or even iron (Table 1).
It is also important to consider the practical implications of the
effect of iron content on sphalerite activation. Under typical plant
conditions, minerals are usually not under an inert environment after
comminution and hence may undergo oxidation. Sphalerite with a
high iron content will oxidise more than sphalerite with low iron
content. In addition minerals in plants are mixed with other minerals
hence galvanic interactions will play a signicant role as these govern
the relative degrees of surface oxidation. It is worth investigating the
effect galvanic interactions (with minerals such as pyrite, galena or
chalcopyrite) have on copper adsorption of high iron sphalerite and
low iron sphalerite. It is envisaged that with a high rest potential
mineral such as pyrite, the reactive nature of high iron sphalerite
should promote more copper adsorption. However, this will be a
function of the degree of surface pre-oxidation.
Pb2+ and Fe2+ present in the mineral slurry as a result of dissolution
from lead and iron containing minerals such as galena and pyrite, and

also due to water recycling, tend to inadvertently activate sphalerite


and cause it to misreport to copper or lead concentrates. Pyrite (if not
sufciently depressed) can also be activated by such contaminant ions.
The exact mechanism by which activation of pyrite by the adsorption
of aqueous lead or iron species occurs is still poorly understood. Ion
exchange and colloidal adsorption mechanisms as a function of pH
have been suggested however the literature lacks sufcient denitive
evidence. Further studies need to be done using mixed mineral
systems and with lead and iron solution concentrations similar to
those encountered in otation circuits. The surface should also be
characterised using the methods suggested further-on in this section.
Determining the exact mechanism of lead and iron activation should
allow the development of techniques and reagents for minimising this
unwanted activation and otation of sphalerite and pyrite into other
concentrates. Investigations should also be made into minimising the
misreporting of sphalerite or pyrite and other gangue materials by
way of entrainment and aggregation through occulation. Reagents
such as sulfoxy species can be used to restore mineral surfaces to
minimise contaminant ion release. Polyphosphates can also be
employed to selectively remove contaminant ions (that have adsorbed
inadvertently) from mineral surfaces. Inert grinding should also be
considered to minimise pulp iron content, in addition to improving
mine water quality.
The study of mixed mineral system is more practical than single
mineral studies and provides a better approximation to actual plant
processing circuits where galvanic interactions between mineral
particles plays a signicant role. From studies conducted with mixed
sphalerite and pyrite it is obvious that a high pH gives better
separation of these two minerals. Pyrite has one of the highest rest
potentials and consequently is usually the cathodic part of any
galvanic couple. In the presence of sphalerite, pyrite becomes covered
with hydroxides (both ferrous and ferric) due to cathodic reduction of
O2 on pyrite surfaces. Pyrite otation is restricted as the hydroxide
species not only obscures sites for copper and xanthate adsorption but
also renders the pyrite surface hydrophilic. In addition sphalerite
being anodic preferentially consumes aqueous copper and hydrophobic cuproussulde and xanthate dominate its surface. Pulp
oxidation potential control is critical for effective galvanic interactions
between pyrite and sphalerite, and O2 conditioning is therefore
needed to maintain sufcient O2 activity. This galvanic interaction can
however be reversed by purging and oating the minerals with an
inert gas like N2. N2 purging reduces pulp potential by decreasing O2
activity, and can be used for reverse otation of pyrite away from
sphalerite (and other anodic minerals) for maximising auriferous
pyrite recovery. The use of an inert gas to reverse the otation
provides conclusive evidence for the existence of galvanic interactions
and associated electrochemical processes. In addition to having the
right potential, using an appropriate copper and xanthate concentration, activation time and particle size should further improve
selectivity. Limits for these parameters will however depend on the
actual ore mineralogy (which changes with time and extent of mining)
and plant operating conditions. New collectors are also emerging that
can offer better selectivity depending on the minerals that need to be
recovered.
Grinding approaches should also be altered to enhance selectivity
and to minimise reagent use. Grinding using steel media uses up O2 in
the pulp causing the pulp oxidation potential to decrease. O2 is
consumed as a result of steel corrosion, which increases when there is
galvanic interaction between the steel media and mineral particles.
This promotes the development of hydrophilic hydroxides on mineral
surfaces (through O2 reduction) while the steel media is oxidised. This
oxidation of steel media is a signicant source of excess iron in
otation pulp. The decreased pulp potential and hydrophilic hydroxides reduce mineral selectivity as galvanic interactions between
mineral particles are reduced. This requires the use of additional O2
and copper/xanthate to restore selectivity. Inert grinding should be

A.P. Chandra, A.R. Gerson / Advances in Colloid and Interface Science 145 (2009) 97110

preferred in practices which require selective otation of an anodic


mineral such as sphalerite.
Selectivity can also be improved through the choice of an
appropriate pyrite depressant such as sulfoxy species (S2 to SO2
3 )
and cyanide. Common sulfoxy reagents are sodium sulte, sulfur
dioxide and sodium metabisulte. They work by enhancing hydroxide
formation on pyrite surfaces through an electrochemical process,
thereby restricting collector adsorption. O2 conditioning may be
needed to restore oxidation potential with sulfoxy addition as the
process may consume O2. The sulfoxy reagents tend to depress pyrite
optimally at high pH and when the delay time between addition of the
sulfoxy reagent and the collector is relatively short. Sodium cyanide
depresses pyrite by formation of stable ironcyanide complexes on
the pyrite surface also through an electrochemical mechanism. It also
displaces collector molecules already adsorbed on the pyrite surface
through an exchange mechanism. Despite the effectiveness of cyanide
as a selective pyrite depressant other more environmentally friendly
alternatives should be found as the cost of environmental and
ecological repair will surpass current prots.
Good water quality should also be maintained for optimal
selectivity and recovery. Mine water is usually impure and recycled
for cost effectiveness however a detrimental quantity of unused
reagents, metal ions, microbiological activity, oxidation products and
precipitates can build up and decrease selectivity and increase reagent
use. Water in the circuits should as such be monitored regularly and
appropriate water treatment techniques employed to maintained
contaminants at an experimentally determined acceptable level.
Efcient sphalerite and pyrite separation can be achieved by use of
high pH, the right oxidation potential and reagent concentration/
conditioning, short time delay between sequential reagent addition,
suitable particle size fraction and minimising pulp contaminants.
Under mixed minerals systems that typies processing plants,
galvanic interactions are rife and electrochemistry is the dominant
process.
One of the very important questions still to be answered is the type
of coppersulde phase formed after copper adsorbs on to the
sphalerite structure. Covellite, chalcocite and metal-decient sulde
type structures have been proposed. There are also suggestions that
Cu(I) has higher than normal 3d9 character and the related sulfur is in
a more positive (or less negative) oxidation state with copper being
bonded to three sulfur atoms. Subsequent oxidation of the sulde to
polysuldes or elemental sulfur also occurs however this may take
place due to zinc dissolution rather than copper adsorption. Such zinc
losses in an actual plant environment will depend on the total time the
mineral is in solution and on conditions such as pH, Eh and the
occurrence of galvanic coupling with other minerals present together
in the slurry. The situation with pyrite is also not clear as structures
similar to those proposed for sphalerite have also been proposed for
pyrite upon copper adsorption.
Copper activation involves surface reactions with copper likely to
penetrate beyond the rst atomic layer, at of sphalerite, within the
activation time. Since surfaces of both sphalerite and pyrite characteristically have defects and low coordination sites along the
cleavage planes, at least some copper on the surface may be in a
different coordination then copper at sites with normal coordination
on the surface. Such structural difference may be limited to within rst
two atomic layers, but copper as we know penetrates several atomic
layers into the sphalerite lattice. In addition there may be preferential
localisation of adsorbed Cu at low coordination sites.
The existence of surface heterogeneity in these minerals and their
relative chemical reactivity are important considerations while
attempting to determine structures on minerals such as sphalerite
and pyrite. This needs techniques with both sufcient surface
sensitivity and a high spatial resolution to characterise the heterogeneous distribution of activation products on (or very near) the
surface environment. As the Cu concentration (near-surface) is

109

relatively low, a synchrotron energy source will be ideal for spatially


resolved measurements. Techniques such as EXAFS, XANES, XRF, XPS
and IR should be considered to obtain information on chemical states,
coordination geometry, species identication and elemental distribution. Other non-synchrotron techniques such as dynamic and static
SIMS can also provide complimentary information on surface species
and its distribution. In addition, theoretical techniques such as
quantum mechanical calculations help validate the existence of
proposed surface structures. Surface heterogeneity of these minerals
(sphalerite and pyrite) however strongly suggests that more than one
coppersulde phase co-exists not only on the surface but also in the
bulk, structures of which may be similar to the ones already proposed.
Another important point to note is that most studies attempt to
identify activated structures on minerals activated under single
mineral system which however providing basic surface chemical
knowledge of the mineral bears no practical relevance. Minerals with
intimate contact with other minerals (typical in-plant situation) will
have galvanic coupling that will enhance Cu adsorption on the anodic
mineral while retarding the same on the cathodic mineral by way of
O2 reduction and subsequent OH production. Investigating and
determining structures on typical mixed mineral systems will be
more meaningful and economically benecial as it can be related
directly to plant practice.
Research studies aimed at improving mineral recovery and
concentrate grades should be performed under conditions typical of
a plant environment. Galvanic interactions occur predominantly and
as such mixed mineral systems should be studied. Reagents and their
concentration should also be typical of plant practice and be applied to
similar mineral compositions as in plants. The water quality and its
redox potential should also be in the similar range as should be the
particle size fraction. While keeping conditions similar to plant
practices is the most prudent approach, deviations to these norms are
also needed to allow the development of new technologies and
approaches. Moreover, it is not possible to entirely replicate plant
environment in the lab and simultaneous plant trials will be a more
practical approach.
Acknowledgments
The support from the University Presidents Scholarship awarded
by University of South Australia to Mr. Anand Chandra is gratefully
acknowledged. This gratitude is further extended towards the
Premiers Science and Research Fund (PSRF) of South Australia, Rio
Tinto and BHP-Billiton for nancial support of the research project
New Information for Minerals Processing.
References
[1] Biswas AK, Davenport WG, editors. Extractive metallurgy of copper. 3rd ed. Oxford:
Pergamon; 1994.
[2] Wills BA, editor. Mineral processing technology: an introduction to the practical
aspects of ore treatment and mineral recovery. 6th ed. Boston: ButterworthHeinemann; 1997.
[3] Hayes PC, editor. Process principles in minerals and materials production. 2nd ed.
Sherwood, Qld.: Hayes Publishing; 1993.
[4] Leppinen JO. Int J Miner Process 1990;30:245.
[5] Finkelstein NP. Int J Miner Process 1997;52:81.
[6] Zhang Q, Xu Z, Bozkurt V, Finch JA. Int J Miner Process 1997;52:187.
[7] Dimitrijevic M, Antonijevic MM, Jankovic Z. Hydrometallurgy 1996;42:377.
[8] Evangelou VP, editor. Pyrite oxidation and its control. New York: CRC Press; 1995.
[9] Woodcock JT, Sparrow GJ, Bruckard WJ, Johnson NW, Dunne R. In: Fuerstenau MC,
Jameson G, Yoon RH, editors. Froth otation: a century of innovations. Colorado:
Society for Mining, Metallurgy, and Exploration, Inc.; 2007.
[10] Natarajan R, Nirdosh I. Int J Miner Process 2006;79:141.
[11] Finch JA, Rao SR, Nesset JE. Iron control in mineral processing, 39th annual meeting
of the Canadian mineral processors 2325 January. Ottawa: Canadian Institute of
Mining, Metallurgy and Petroleum; 2007.
[12] Finkelstein NP, Poling GW. Miner Sci Eng 1977;9:177.
[13] Gerson AR, Lange AG, Prince KE, Smart RSC. Appl Surf Sci 1999;137:207.
[14] Fornasiero D, Ralston J. Int J Miner Process 2006;78:231.
[15] Popov SR, Vucinic DR. Int J Miner Process 1990;30:229.

110

A.P. Chandra, A.R. Gerson / Advances in Colloid and Interface Science 145 (2009) 97110

[16] Sutherland KL, Wark IW, editors. Principles of otation. Melbourne: Australasian
Institute of Minning and Metallurgy; 1955.
[17] Pattrick RAD, England KER, Charnock JM, Mosselmans JFW. Int J Miner Process
1999;55:247.
[18] Prestidge CA, Skinner WM, Ralston J, Smart RSC. Appl Surf Sci 1997;108:333.
[19] Kartio IJ, Basilio CI, Yoon RH. Langmuir 1998;14:5274.
[20] Buckley AN, Woods R, Wouterlood HJ. Int J Miner Process 1989;26:29.
[21] Buckley AN, Skinner WM, Harmer SL, Pring A, Lamb RN, Fan L, et al. Can J Chem
2007;85:767.
[22] Vinals J, Fuentes G, Hernandez MC, Herreros O. Hydrometallurgy 2004;75:177.
[23] Chen Z, Yoon RH. Int J Miner Process 2000;58:57.
[24] Gerson AR, Cookson DJ, Prince KC. In: Riviere JC, Myhra S, editors. Handbook of
surface and interface analysis: methods for problem solving, 2nd Edition; 2008.
Chapter 10.
[25] Porento M, Hirva P. Surf Sci 2005;576:98.
[26] Popov SR, Vucinic DR. Colloids Surf 1990;47:81.
[27] Solecki J, Komosa A, Szczypa J. Int J Miner Process 1979;6:221.
[28] Bacaksiz E, Dzhafarov TD, Novruzov VD, Ozturk K, Tomakin M, Kucukomeroglu T,
et al. Phys Status Solidi. A Appl Res 2004;201:2948.
[29] Buckley AN, Woods R. Appl Surf Sci 1987;27:437.
[30] Lascelles D, Sui CC, Finch JA, Butler IS. Colloids surf, A Physicochem Eng Asp
2001;186:163.
[31] Weisener C, Gerson A. Miner Eng 2000;13:1329.
[32] Skinner WM, Prestidge CA, Smart RSC. Surf Interface Anal 1996;24:620.
[33] Gigowski B, Vogg A, Wierer K, Dobias B. Int J Miner Process 1991;33:103.
[34] Zhang Q, Rao SR, Finch JA. Colloids Surf 1992;66:81.
[35] Rao SR, editor. Surface chemistry of froth otation. Fundamentals, 2nd edition.
New York: Kluwer Academic/Plenum Publishers; 2004.
[36] Mirnezami M, Restrepo L, Finch JA. J Colloid Interface Sci 2003;259:36.
[37] Boulton A, Fornasiero D, Ralston J. Int J Miner Process 2003;70:205.
[38] Boulton A, Fornasiero D, Ralston J. Miner Eng 2005;18:1120.
[39] Duarte ACP, Grano SR. Miner Eng 2007;20:766.
[40] He S, Fornasiero D, Skinner W. Miner Eng 2005;18:1208.
[41] He S, Skinner W, Fornasiero D. Int J Miner Process 2006;80:169.
[42] Huang G, Grano S. Miner Eng 2005;18:1152.
[43] Khmeleva TN, Skinner W, Beattie DA. Int J Miner Process 2005;76:43.
[44] Wang XH, Forssberg KSE. Miner Eng 1996;9:527.
[45] Dana JD, Hurlbut CS, Klein C, editors. Manual of mineralogy (after James D. Dana).
19th ed. New York: Wiley; 1977.
[46] Harmer SL, Goncharova LV, Kolarova R, Lennard WN, Munoz-Marquez MA,
Mitchell IV, et al. Surf Sci 2007;601:352.
[47] Wenk H-R, Bulakh AG, editors. Minerals: their constitution and origin. Cambridge;
New York: Cambridge University Press; 2004.
[48] Harmer SL, Mierczynska-Vasilev A, Beattie DA, Shapter JG. Miner Eng in press,
doi:10.1016/j.mineng.2008.02.014.
[49] Szczypa J, Solecki J, Komosa A. Int J Miner Process 1980;7:151.
[50] Weisener CG, Smart RSC, Gerson AR. Int J Miner Process 2004;74:239.
[51] Morey MS, Grano SR, Ralston J, Prestidge CA, Verity B. Miner Eng 2001;14:1009.
[52] Rashchi F, Sui C, Finch JA. Int J Miner Process 2002;67:43.
[53] Sui CC, Lee D, Casuge A, Finch JA. Miner Metall Process 1999;16:53.

[54] Yeomans T. Copper concentrate quality improvements at Myra Falls. In: Orford I,
editor. 40th Annual meeting of the Canadian mineral processors 2224 January.
Ottawa: Canadian Institute of Mining, Metallurgy and Petroleum; 2008.
[55] Basilio CI, Kartio IJ, Yoon RH. Miner Eng 1996;9:869.
[56] Pattrick RAD, Charnock JM, England KER, Mosselmans JFW, Wright K. Miner Eng
1998;11:1025.
[57] O'Dea AR, Prince KE, Smart RSC, Gerson AR. Int J Miner Process 2001;61:121.
[58] Finkelstein NP. Int J Miner Process 1999;55:283.
[59] Tong X, Song S, He J, Rao F, Lopez-Valdivieso A. Miner Eng 2007;20:259.
[60] Majima H. Can Metall Q 1969:269.
[61] Zielinski PA, Larson KA, Stradling AW. Miner Eng 2000;13:357.
[62] Lascelles D, Finch JA. Miner Eng 2002;15:567.
[63] Wong G, Lascelles D, Finch JA. Miner Eng 2002;15:573.
[64] Miller JD, Kappes R, Simmons GL, LeVier KM. Miner Eng 2006;19:659.
[65] Pecina ET, Uribe A, Nava F, Finch JA. Miner Eng 2006;19:172.
[66] Valdivieso LA, Sanchez Lopez AA, Song S. Int J Miner Process 2005;77:154.
[67] Weisener C, Gerson A. Surf Interface Anal 2000;30:454.
[68] Hicyilmaz C, Emre Altun N, Ekmekci Z, Gokagac G. Miner Eng 2004;17:879.
[69] Shen WZ, Fornasiero D, Ralston J. Int J Miner Process 2001;63:17.
[70] Dichmann TK, Finch JA. Miner Eng 2001;14:217.
[71] Shen WZ, Fornasiero D, Ralston J. Miner Eng 1998;11:145.
[72] Rao SR, editor. Surface chemistry of froth otation. Reagents and Mechanisms, 2nd
edition, New York: Kluwer Academic/Plenum Publishers; 2004.
[73] Miller JD, Du Plessis R, Kotylar DG, Zhu X, Simmons GL. Int J Miner Process
2002;67:1.
[74] Hart B, Biesinger M, Smart RSC. Miner Eng 2006;19:790.
[75] Pease JD, Curry DC, Young MF. Miner Eng 2006;19:831.
[76] Rashchi F, Finch JA. Colloids Surf, A Physicochem Eng Asp 2006;276:87.
[77] Kocabag D, Gler T. Miner Eng 2007;20:1246.
[78] Fuerstenau MC, Chander S, Woods R. In: Fuerstenau MC, Jameson G, Yoon RH,
editors. Froth otation: a century of innovation. Colorado: Society for Mining,
Metallurgy, and Exploration, Inc.; 2007.
[79] Valdivieso LA, Celedn Cervantes T, Song S, Robledo Cabrera A, Laskowski JS. Miner
Eng 2004;17:1001.
[80] Rao KH, Forssberg KSE. In: Fuerstenau MC, Jameson G, Yoon RH, editors. Froth
otation: a century of innovation. Colorado: Society for Mining, Metallurgy, and
Exploration, Inc.; 2007.
[81] Rao SR, Nesset JE, Finch JA. Activation of sphalerite by Cu ion produced by cyanide
action on chalcopyrite. In: Khosla NK, Jadhav GN, editors. Proceedings of the
International Seminar on Mineral Processing Technology (MPT) 2007. New Delhi:
Allied Publishers; 2007.
[82] Levay G, Smart RSC, Skinner WM. J S Afr Inst Min Metall 2001;101:69.
[83] Rao SR, Finch JA. Miner Eng 1989;2:65.
[84] Ozkan SG, Acar A. Water Res 2004;38:1773.
[85] Espinosa-Gomez R, Finch JA, Laplante AR. Colloids Surf 1987;26:333.
[86] Seke MD, Pistorius PC. Miner Eng 2006;19:1.
[87] Stn P, Parvinen P, Miettinen M, Luukkanen S, Kaskiniemi V, Aaltonen J. Miner Eng
2003;16:229.
[88] El-Shall H, Zhang P, Snow R. Miner Metall Process 1996;13:135.
[89] Rodrigues RT, Rubio J. Int J Miner Process 2007;82:1.

You might also like