You are on page 1of 18

G Model

JIEC 3218 No. of Pages 18

Journal of Industrial and Engineering Chemistry xxx (2016) xxxxxx

Contents lists available at ScienceDirect

Journal of Industrial and Engineering Chemistry


journal homepage: www.elsevier.com/locate/jiec

Review

A review on electrochemical behavior of pyrite in the froth otation


process
Hossein Moslemia,* , Mahdi Gharabaghib
a
b

Department of Mining & Metallurgical Engineering, Amirkabir University of Technology, Hafez St., Tehran, Iran
School of Mining Engineering, College of Engineering, University of Tehran, Tehran, Iran

A R T I C L E I N F O

Article history:
Received 29 October 2016
Received in revised form 10 December 2016
Accepted 18 December 2016
Available online xxx
Keywords:
Mineral processing
Pyrite
Flotation
Electrochemical potential
Surface chemistry

A B S T R A C T

Metal suldes are usually semiconductor and cause electrochemical reactions. This phenomenon plays
an important role in sulde otation. Pyrite as the most abundant sulde mineral is often associated with
valuable sulde minerals, coal and gold. It is very important to study its electrochemical behavior in the
otation process. This review focuses on researches carried out over the past several decades that have
studied electrochemical processes associated with pyrite occurring during otation. The mechanism of
processes such as oxidation, activation, depression, and interactions of activated and non-activated
surfaces with collectors as well as factors affecting them are described. Moreover, the effect of
electrochemical conditions during grinding on the otation process is also discussed. It has been found
that moderately oxidizing conditions are favorable for collector-less otation of pyrite while strongly
reducing or oxidizing potentials lead to its depression. Increasing the electrochemical potential not only
has a deleterious effect on the activation of pyrite by copper, but also facilitates its depression by
depressants. In the case of the adsorption of xanthate whether on activated or non-activated surfaces, a
great increase or decrease in the potential has adverse effects and it is necessary to optimize the
electrochemical conditions. Various factors such as pH, solid percentage, particle size distribution,
otation time, type and concentration of reagents and oxygen content as well as grinding conditions can
affect the intensity of these electrochemical interactions. It is proposed that further researches using
advanced chemical analysis techniques are needed to understand the electrochemical processes involved
in otation systems.
2016 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights
reserved.

Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Oxidation of pyrite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Effect of oxidation on pyrite oatability . . . . . . . . . . . . . . . . .
Effect of pyrite origin on the oatability due to oxidation
Effect of oxidation on pyrite depression . . . . . . . . . . . . . . . . .
Effect of pH on pyrite depression . . . . . . . . . . . . . . . . . . .
Effect of dissolved oxygen content on pyrite depression .
Removal of depressing effect of oxidation . . . . . . . . . . . .
Pyrite activation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Pyrite activation by copper . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Pyrite activation by lead . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Collector adsorption on pyrite surface . . . . . . . . . . . . . . . . . . . . . .
Collector adsorption on non-activated surface . . . . . . . . . . . .

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

* Corresponding author.
E-mail addresses: hoseinn_moslemi@yahoo.com, h.moslemi@aut.ac.ir
(H. Moslemi), gharabaghi@ut.ac.ir, m.gharabaghi@gmail.com (M. Gharabaghi).
http://dx.doi.org/10.1016/j.jiec.2016.12.012
1226-086X/ 2016 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights reserved.

Please cite this article in press as: H. Moslemi, M. Gharabaghi, A review on electrochemical behavior of pyrite in the froth otation process, J.
Ind. Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.12.012

00
00
00
00
00
00
00
00
00
00
00
00
00

G Model
JIEC 3218 No. of Pages 18

H. Moslemi, M. Gharabaghi / Journal of Industrial and Engineering Chemistry xxx (2016) xxxxxx

Effect of pH on collector adsorption . . . . . . . . . . . . . . . . . . . . . . . . .


Effect of oxygen content on collector adsorption . . . . . . . . . . . . . . .
Effect of pyrite surface oxidation products on collector adsorption
Effect of pyrite type and origin on the collector adsorption . . . . . .
Collector adsorption on activated surface . . . . . . . . . . . . . . . . . . . . . . . .
Collector adsorption on copper activated surface . . . . . . . . . . . . . .
Collector adsorption on lead activated surface . . . . . . . . . . . . . . . . .
Depressants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Inorganic depressants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Organic depressants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Effect of grinding environment on pyrite electrochemical properties . . . . . .
Effect of grinding media type . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Effect of oxygen content in grinding system . . . . . . . . . . . . . . . . . . . . . .
Effect of water quality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Effect of reactions occurred in the grinding system . . . . . . . . . . . . . . . .
Summary and recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Introduction
Demand for base metals has been increasing continuously due
to the development of industries worldwide; therefore, it is
necessary to improve processing of sulde ores which are the
major sources of base metals. Pyrite (FeS2, iron disulde), which is
the most common sulde mineral on the Earth [15], is the main
gangue of sulde ores [16]. The pyrite presence in the concentrate
of valuable minerals leads to a decrease in grade and an increase in
sulfur and iron contamination resulting in an increase in smelting
costs [1,4,7]. Moreover, if pyrite contains potentially hazardous
elements such as arsenic, hazardous dusts and fumes will be
produced during the smelting process [812]. Pyrite is also found
in coal deposits [13]. The burning process of coal containing a high
content of pyrite leads to the atmospheric emission of sulfur oxides
[1419] which are the major cause of acid rain [1421]. The
primary aim of the processing of sulde ores and coal is to remove
the pyrite content [22]. However, occasionally, pyrite may contain
signicant amounts of valuable metals such as gold [2330], and
thus, it may be concentrated to gain these valuable metals [23].
Froth otation is widely used in mineral industry to selectively
separate minerals from each other [31]. Annually, more than
109 tons of various materials are processed by this method
worldwide [32]. The process strongly depends on the physicochemical surface properties of minerals [14,33] and is controlled by
modifying these properties through addition of otation reagents
[14].
Since most metal suldes are semiconductor [3437], various
electrochemical reactions occur in the sulde mineral otation
system. Extensive researches have shown that there is a strong
correlation between otation of pyrite and electrochemical
reactions [38]. Electrochemical mechanisms are often known to
be responsible for various phenomena, occurring in a otation
system containing pyrite, such as changes in pyrite surface
chemistry due to oxidation, the interaction of pyrite with other
components, adsorption of collector and precipitation of metals on
the surface [39]. The most important factor which affects
electrochemical processes is the electrochemical potential of
mineral/solution interface [40,41]. This potential is a mixed
potential in which the rates of anodic and cathodic reactions,
occurring on the mineral surface, are exactly equal to each other
[40,4244]. The electrochemical potential controls formation of
not only surface species responsible for the otation of the mineral
(such as polysuldes, elemental sulfur and xanthate), but also
surface species responsible for its depression (such as ferric oxide/
hydroxide and sulfate) [45].

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

00
00
00
00
00
00
00
00
00
00
00
00
00
00
00
00
00

Since electrochemical potential is known as the key factor in


pyrite otation process, the present study was performed to
provide a comprehensive review of the effects of electrochemical
conditions on pyrite surface chemistry and also electrochemical
processes occurring on the pyrite surface in the otation system.
Oxidation of pyrite and its effects on the otation, activation of
pyrite, depression of pyrite, are also discussed in detail in this
review. In addition, interactions between pyrite and collectors and
interactions between pyrite and grinding media with emphasis on
the electrochemical reactions were focused. There are some
researches which investigate various interactions in the pyrite
otation, so this review is intended to provide further insights into
the pyrite otation process and to address important points for the
control of the process and its optimization in order to enhance the
product quality and to reduce process costs.
Oxidation of pyrite
Pyrite is oxidized in aqueous solutions through an electrochemical mechanism. The oxidation rate is inuenced by various
factors such as solution pH, solution electrochemical potential,
oxidant type and concentration, particle size, temperature and
agitation speed [4653]. Different samples of pyrite exhibit
different electrochemical reactivity [23] due to various reasons
such as differences in iron to sulfur ratio, crystal structure and
surface morphology [5456].
If pyrite is completely oxidized, ferrous iron and sulfate ions are
produced [57,58]. The complete oxidation of pyrite occurs
according to Reaction (1) [5961].
FeS2 + 3.5O2 + H2O ! Fe2+ + 2SO42 + 2H+

(1)

Experimental observations indicate that the pyrite oxidation


does not completely occur and elemental sulfur is also produced in
addition to the ferrous iron and the sulfate ions as a result of
incomplete oxidation (Reaction (2)) [6269].
FeS2 + 2.9O2 + 0.6H2O ! Fe2+ + 0.4S0 + 1.6SO42 + 1.2H+

(2)

The ferrous iron produced by Reactions (1) and (2) undergoes


Reactions (3) and (4) and forms ferric hydroxide [70,71].
Fe2+ + 0.25O2 + H+ ! Fe3+ + 0.5H2O

(3)

Fe3+ + 3H2O ! Fe(OH)3 + 3H+

(4)

The formed ferric hydroxide can be precipitated as a coating on


the pyrite surface of and inhibits further oxidation of pyrite. This

Please cite this article in press as: H. Moslemi, M. Gharabaghi, A review on electrochemical behavior of pyrite in the froth otation process, J.
Ind. Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.12.012

G Model
JIEC 3218 No. of Pages 18

H. Moslemi, M. Gharabaghi / Journal of Industrial and Engineering Chemistry xxx (2016) xxxxxx

phenomenon occurs at higher rates with increasing solution pH


[72,73].
The electrochemical oxidation of pyrite, producing ferrous and
sulfate ions, involves removal of seven electrons from each sulfur
atom to oxidize S1 in the pyrite, to S6+ in the sulfate ions. Since
only one or at most two electrons can be removed at any one time,
the oxidation process is quite complex, involving several key
intermediate electrochemical reactions with the formation of
various sulfoxy species [74]. The presence and abundance of these
species on the surface depends strongly on the electrochemical
conditions [4].
The most important step in the pyrite oxidation process is the
breaking of strong bond between sulfur and iron atoms, and
formation of thiosulfate ion as intermediate product (Reaction (5))
[51].
FeS2 + 1.5O2 ! Fe2+ + S2O32

(5)

The produced thiosulfate ion is unstable, especially in acidic


conditions [51], and decomposes to elemental sulfur and
tetrathionate ion through Reaction (6) [69].
S2O32 + 1.2H+ ! 0.4S0 + 0.4S4O62 + 0.6H2O

(6)

The formed tetrathionate ion, which is also an unstable


intermediate product [51], is further oxidized to form sulfate
ion as a nal product according to the Reaction (7) [69].
S4O62 + 3.5O2 + 3H2O ! 4SO42 + 6H+

(7)

Despite extensive studies conducted on the oxidation of pyrite,


there is little agreement on the details of oxidation processes
particularly surface composition of the oxidized pyrite [55]. Using
spectroscopic surface and electrochemical analysis, some
researchers have found that mild oxidation resulted in the
formation of an iron-decient surface layer on pyrite [65
67,75,76] while a number of researchers believe the oxidation
product of pyrite is iron polysulde [7780]. On the other hand,
using linear sweep voltammetry, it was observed that a monolayer
or multiple layers of sulfur (depending on solution pH) were
formed on the surface of oxidized pyrite [57]. It was also found that
sulfur forms vary from disulde ion (in pyrite) on the internal
surface to elemental sulfur (which is suitable for electrochemical
reaction) on the outer surface [81].
In general, most of these studies have shown that the
concentration of sulfur in a pyrite oxidized layer increases from
the internal to the outer surface while the concentration of iron
decreases in this direction [81]. However, under strongly oxidizing
conditions where sulfate is the main oxidation product [57,62
64,82,83], it has been observed that the sulfur concentration on the
surface is reduced relative to iron [54].
Fig. 1 shows the Eh-pH diagram for FeS2-H2O system at 25  C.
The elds of thermodynamic stability of various species of iron and
sulfur in terms of pH and electrochemical potential (Eh) are
presented in this diagram [84].
Effect of oxidation on pyrite oatability
Pyrite is naturally oatable under specic conditions and a
proportion of it can be recovered without use of any otation
reagent [31,45,85]. Small changes in the surface chemistry of
pyrite may signicantly affect its oatability [86]. These changes
can be due to the anodic oxidation of the mineral surface and
the cathodic reduction of oxygen dissolved in the pulp [62].
Electrochemical properties such as redox potentials of the system
and rest potentials of the mineral are used to describe the changes
in the surface composition of pyrite [87,88]. Some researchers
have attempted to measure the variation in the otation pulp

Fig. 1. Eh-pH diagram for FeS2H2O system at 25  C and for 105 M dissolved
species [84].

potential and to study their effects on the pyrite otation


recovery.
Tao et al. [89] conducted studies to measure the pyrite otation
recovery in a micro-otation electrochemical cell at pH 4.6 and pH
9.2. The collector-less otation obtained results recovery as a
function of pulp potential- showed that the recovery began to
increase at potential values close to 0 and 100 mV (SHE1 ) at pH
values 4.6 and 9.2, respectively.
In another study, they observed that, in the absence of collector
at pH 9.2, maximum otation of pyrite occurred at potentials
slightly above 0.28 V (SHE) [54].
Results of collector-less otation tests performed by Yoon et al.
[90] showed that the pyrite otation occurred in potential ranges
of 0.1 to 0.8 V and 0.3 to 0.4 V (SHE) at pH 4.6 and 9.2, respectively.
The results also showed that the otation recovery at pH 4.6 was
considerably higher than recovery at pH 9.2.
Hicyilmaz et al. [49] measured pyrite contact angle in the
absence of collector in the potential range of 505 to +595 mV
(SHE) and at different pH values. The maximum measured contact
angle was 25 at the potential of +400 mV (SHE) at pH 4.67. The
variation trend of the contact angle as a function of potential was
similar at all pH values.
Guler et al. [91] investigated oatability of pyrite in the
collector-less conditions in a wide potential range involving
reducing (400 mV, 100 mV (SHE)), slightly oxidizing
(+200 mV (SHE), around open circuit potential) and highly
oxidizing (+500 mV, +800 mV (SHE)) potentials in borate buffer
solution at pH 9.2. Results showed that the recovery increased with
potential increasing and reached its maximum (approximately
18%) at about 200250 mV (SHE) and then decreased at more
oxidizing potentials.
The relationship between the pulp electrochemical potential
and pyrite otation recovery in the collector-less conditions from
various investigations are summarized in Table 1.
The results of these electrochemical studies on pyrite otation
systems clearly show that the otation recovery in the collectorless conditions is a function of the mineral/solution interface

SHE: standard hydrogen electrode.

Please cite this article in press as: H. Moslemi, M. Gharabaghi, A review on electrochemical behavior of pyrite in the froth otation process, J.
Ind. Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.12.012

G Model
JIEC 3218 No. of Pages 18

H. Moslemi, M. Gharabaghi / Journal of Industrial and Engineering Chemistry xxx (2016) xxxxxx

Table 1
Relationship between pulp electrochemical potential and pyrite otation recovery
in collector-less conditions.
Pulp potential
(mV vs. SHE)

Pulp pH

Reference

Increased
Was observed
Maximum
Increased
Maximum
Was observed
Maximum

[89]
[90]
[49]
[89]
[54]
[90]
[91]

Alkaline

Acidic
0
100 to 800
400
100
280
300 to 400
200 to 250

Recovery

4.6
4.6
4.67
9.2
9.2
9.2
9.2

electrochemical potential [92]. Cyclic voltammetry studies of


freshly created pyrite surfaces indicate that the potentials at which,
pyrite begins to oat correspond to the potentials at which, the
freshly created pyrite surfaces begin to oxidize [89]. Therefore, it
can be concluded that the initial oxidation of pyrite occurs at
potentials slightly more positive than the stable potential, results
in an increase in pyrite oatability [39]. (The stable potential is the
potential at which, the surface of pyrite is neither oxidized nor
reduced) [54].
Tao et al. [89] using XPS2 and electrochemical studies proposed
that metal-decient suldes (Fe1xS2, x < 1) and iron polysuldes
(FeSn, n > 2) are the sulfur oxidation product(s) and are responsible
for the pyrite otation. The reactions producing these species can
be represented by Reactions (8) and (9).
FeS2 + 3x(OH) = Fe1xS2 + xFe(OH)3 + 3xe

(8)

nFeS2 + 3(n  2)OH = 2FeSn + (n  2)Fe(OH)3 + 3(n  2)e

(9)

XPS studies conducted by Buckley and Woods [67] showed that


moderate oxidation of pyrite released iron preferentially from
normal lattice sites and resulted in creating a metal-decient
surface through formation of soluble iron compounds. Since the
oatability of pyrite is inuenced by the relative amount of
hydrophobic sulfur to insoluble oxide/hydroxides species present
on the surface [89], a decrease in the amount of iron hydroxide
results in an increase in the oatability [54]. The role of the
polysuldes and the metal-decient suldes as the oxidation
products responsible for the hydrophobicity of pyrite has been
conrmed by researchers [7880,9396].
In addition to the polysuldes and the metal-decient suldes,
it has been found that the elemental sulfur can also be responsible
for the collector-less oatability [31,57,67,97,98]. Ekmekci and
Demirel [63] obtained a pyrite recovery over 90% after 10 min of
otation at pH 4.6. Their studies showed that this high recovery
was due to the formation of the hydrophobic elemental sulfur on
the surface. The elemental sulfur is believed to be a critical factor in
collector-less otation [38,88]. Researchers have reported that the
elemental sulfur is formed on the pyrite surface by the oxidation of
pyrite and its decomposition (Reaction (10)) [63],
FeS2 = Fe2+ + 2S0 + 2e

(10)


or by the oxidation of HS ions on the pyrite surface (Reaction (11))


[88,99,100]:
HS = S0 + H+ + 2e


(11)

The HS ion may be produced by the cathodic reduction of the


oxidation products on the pyrite surface [47]. This ion is expected
to be formed at neutral and alkaline pH values [99]. Based on the

XPS: X-ray photoelectron spectroscopy.

signicant difference between the recovery values of 5 min (50%)


and 10 min (93%) of otation, Ekmekci and Demirel [63] also
suggested that in this time interval, the hydrophobic surface
species are slowly produced by the electrochemical reaction.
Effect of pyrite origin on the oatability due to oxidation
There are differences between mineral-pyrites and coal-pyrites
in the oatability and electrochemical properties. It was reported
that coal-pyrite could be oated over a wide range of pH values
while mineral-pyrite could only be oated at acidic pH. It was also
observed that in the absence of xanthate collectors and at a pH
above 6.5, coal-pyrite exhibited signicantly higher oatability
than mineral-pyrite [90]. In contrast, Yoon et al. [90] observed that
otation recovery of coal-pyrite was lower than mineral-pyrite.
Electrochemical studies showed that coal-pyrite produces a higher
concentration of oxide species on its surface. This can be due to the
fact that coal-pyrite is poorly crystalline and has a higher surface
area. Therefore, its oxidation occurs faster than mineral pyrite
[101103] and its metastable hydrophobic sulfur-rich species are
more readily transformed to the hydrophilic oxidation species
which decreases its oatability [101,102].
In addition to the oatability differences between mineralpyrite and coal-pyrite, there are also differences in the oatability
of pyrites from different geographical locations. For example,
Gebhart et al. [104] observed that the pyrite collector-less otation
was maximum at potentials of approximately +85 mV (Ag/AgCl),
whereas Chandurya et al. [105] obtained maximum pyrite otation
at 700 mV (Ag/AgCl) without collector at pH 9.2. This difference
in the pyrite otation behavior may be due to the presence of
lattice defects in the pyrite bulk due to the lattice substitution of Fe
and S by impurities such as As, Co, Ni, Pb, Cu, Se, Zn, Au and Ag
which inuences the electrochemical properties of pyrite [106
112]. Xian et al. [106] conducted otation studies on four types of
pyrite, including perfect pyrite, As-substituted pyrite, Co-substituted pyrite and intercrystalline Au pyrite. In the perfect pyrite
bulk, each S atom was coordinated to three Fe atoms and one S
atom in a tetrahedral conguration, whereas in the bulk of Assubstituted pyrite an As atom substituted for one of S atoms, in the
bulk of Co-substituted pyrite a Co atom substituted for one Fe atom
and in the bulk of intercrystalline Au pyrite an Au atom lled in the
interstitial positions. Fig. 2 shows the lattice structures of the four
types of pyrite proposed by Xian et al. [106].
The otation results showed that the oatability of cosubstituted and intercrystalline Au pyrites increased with pulp
aeration time, whereas oatability of pure and As-substituted
pyrites decreased with pulp aeration time. Using electronic
structure and band structure studies of the four types of pyrites,
Xian et al. [106] observed that the stability of the pyrites was
inuenced by the impurity defects and the electronic structures.
The stability of the pyrites increased in the following order:
As-substituted < perfect < Co-substituted < intercrystalline Au.
Therefore, it was concluded that the observed difference in the
otation behavior was due to the difference in the stability of the
pyrites and their oxidation intensity. The effect of the presence of
such impurities in the pyrite lattice on the oxidation process has
also been conrmed by other researchers [113115]. It has been
demonstrated that the electrochemical properties and the otation
behavior of pyrite is also signicantly inuenced by vacancy
defects due to loss of anions or cations in the pyrite lattice [116].
Effect of oxidation on pyrite depression
Several investigations have been conducted in order to evaluate
the inuence of the electrochemical potential on the surface
properties of pyrite leading to its depression. Hicyilmaz et al. [49]
measured contact angles of pyrite in the absence of collector in a

Please cite this article in press as: H. Moslemi, M. Gharabaghi, A review on electrochemical behavior of pyrite in the froth otation process, J.
Ind. Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.12.012

G Model
JIEC 3218 No. of Pages 18

H. Moslemi, M. Gharabaghi / Journal of Industrial and Engineering Chemistry xxx (2016) xxxxxx

Fig. 2. Lattice structures of four types of pyrite; perfect (a), As-substituted (b), Co-substituted (c) and intercrystalline Au (d) [106].

potential range of 505 to +595 mV (SHE) and observed that


increasing pulp potential from +400 to +595 mV (SHE), decreased
contact angles. Similar results were also obtained in other study
[91]. The decrease in the contact angles and oatability of pyrite
with increasing pulp potential was found to be due to the evolution
and precipitation of the hydrophilic species such as iron hydroxide
at high potentials [63,117,118]. The performed studies have shown
that since ferrous ions released by oxidation of pyrite have low
stability, these ions are then oxidized to the ferric oxy-hydroxide
species at highly oxidizing potentials. These nal products of
oxidation are hydrophilic, and thus decrease the pyrite oatability
and eventually cause its depression [91]. Another reason for the
loss of the hydrophobicity at high potentials can also be the
oxidation of the hydrophobic elemental sulfur into the hydrophilic
species such as sulfate [119121] according to following reaction
[49]:
S + 4H2O = SO42 + 8H+ + 6e

(12)

The percentage of various elements on the pyrite surface at


electrochemical potential values of 60 and +130 mV (SHE) was
measured by XPS surface analysis [45]. The results showed that at
positive potential values, percentage of the surface oxidation
species (%O) was more than the sulfur species while the opposite
was true at the negative potential value. The results of zeta
potential measurements and XPS studies also supported this fact
that the concentration of these surface oxidation species which
were formed on the pyrite surface increases with an increase in the
electrochemical potential [122]. The adverse effects of the

extensive oxidation products such as SO32, S2O32 and SO42 and


the iron hydroxide species on the pyrite oatability have also been
reported by various researchers [22,58,63,93,99,123,124].
It is possible to minimize the oatability of pyrite by modifying
the pulp potential at an appropriate value in order to oxidize the
hydrophobic sulfur species to sulfate [54]. On this basis, oxidation
at the anode has been considered as the key step for electrochemical desulfurization of coal [125]. On the other hand, it was found
that reduced pyrite is more hydrophobic than pyrite. Therefore, in
the case of coal, grinding and otation under moderate reducing
potentials have also been proposed as an effective method for
preventing pyrite oxidation and improving coal desulfurization
[89,126]. It was reported that galvanic coupling of pyrite with
active metals such as manganese, zinc and aluminum can be a
practical technique for creating a reducing environment. Using this
technique, decreasing the potential of pyrite to values that were
negative enough to inhibit the oxidation of pyrite and to reduce the
hydrophobic oxidation products that were already present on the
surface was successfully performed [90]. The result of otation
experiments performed on coal samples conditioned with an
active metal powder obviously showed that the separation
efciency was signicantly improved by the galvanic coupling.
For example, the pyrite rejection increased from 45% in the absence
of manganese to 75% in the presence of manganese.
Effect of pH on pyrite depression
It has been found that the pyrite otation recovery decreases
with an increase in pH [54,63,89,122,127130]. This inverse

Please cite this article in press as: H. Moslemi, M. Gharabaghi, A review on electrochemical behavior of pyrite in the froth otation process, J.
Ind. Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.12.012

G Model
JIEC 3218 No. of Pages 18

H. Moslemi, M. Gharabaghi / Journal of Industrial and Engineering Chemistry xxx (2016) xxxxxx

relation between the recovery otation and pH is associated with


the relative abundance of the hydrophilic hydroxide to the
hydrophobic sulfur species [97]. The relative abundance of the
hydrophilic to the hydrophobic species is expected to be largely
dependent on the solubility of the oxidation products [54]. It has
been found that as the pyrite surface is oxidized, ferrous ions
move from the mineral lattice to the surface due to their high
mobility. In the alkaline environments, released ferrous ions
undergo further reactions to form ferrous hydroxides. The
produced ferrous hydroxides are then oxidized to the ferric
hydroxides and precipitated on the pyrite surface [63,83,131]. The
formation rate and stability of the hydrophilic products of
oxidation increase with increasing pH [63,72,132134]. The
hydrophobic species such as sulfur are also formed by the
pyrite oxidation, but they are covered by the hydrophilic species
such as ferric hydroxides [131]. The coverage of the pyrite
surface by the stable iron oxy-hydroxide species during the
oxidation process has also been conrmed by various researchers
[49,50,72,135,136].
Measurement of pyrite contact angles as a function of pH
showed that the highest values were obtained at slightly acidic
pH [49]. It was also found that with increasing pH and passing
from acidic to neutral and alkaline conditions, hydrophobicity of
the pyrite surface signicantly decreased. Peng et al. [122]
conducted otation studies in the absence of collector at pH 9 and
in the potentials of 185, 10 and +260 mV (SHE) and obtained
very low recovery values (around 8 wt. %) at the completion of
8 min of otation. They concluded that in the alkaline conditions,
the electrochemical potential has no signicant effect on the
otation recovery. As it was stated previously, at alkaline pH
values, formed iron oxy-hydroxide species are electrochemically
very stable and cause extensive passivation of the pyrite surface.
This surface passivation in the alkaline solutions causes the
mineral to exhibit no response, or a negligible response, to the
changes in the pulp potential [63,137,138]. Cyclic voltammetry
studies by Hicyilmaz et al. [49] on pyrite samples in the strongly
alkaline solution (i.e. pH 11) also supported this nding, where no
apparent cathodic or anodic peaks were observed in the obtained
voltammograms.
Effect of dissolved oxygen content on pyrite depression
The effect of dissolved oxygen content in the pulp on the
variation of the pulp potential and the formation of the
hydrophilic species on the pyrite surface has also been
investigated by various researchers. Owusu et al. [139] measured
pH, pulp potential (Eh) and dissolved oxygen content of pyrite
pulp after various aeration times of the pulp. Initially, the pulp
was purged with air for 3 min and the dissolved oxygen content
was then measured with time. They observed that the dissolved
oxygen concentration of the pulp reduced from 5.5 to 0.06 mg/
dm3 during 5 min. This showed that the oxygen dissolved in the
pulp was nearly completely consumed by pyrite. However, with
increasing aeration time, it was observed that the residual
dissolved oxygen content increased to 5.21 mg/dm3. On the other
hand, it was also observed that with increasing the dissolved
oxygen content of the pulp, the pulp pH decreased considerably
and the pulp Eh changed from a reducing to more oxidizing value.
These changes in pH and Eh of the pulp were found to be favorable
for the pyrite oxidation and the oxygen consumption. Therefore, it
was concluded that by increasing aeration time, formation of iron
oxy-hydroxide species and also coating the surface with these
species increase. Then, however, because of the coverage of most
of the pyrite surface by these species, a signicant decrease in the
surface reactivity with oxygen occurs. In other words, with a
further increase in aeration time, the pyrite surface is more
passivated [139].

Removal of depressing effect of oxidation


The adverse effects of the extensive oxidation products on the
pyrite otation can be eliminated or reduced through various
methods. It has been reported that reducing agents are able to
restore the potential to the optimum level and to reduce the
oxidation products [140]. These products which formed on the
surface, however, may be not completely removed by this method
because of the irreversibility of their reactions [53,140,141].
EDTA3 solutions have also been found to be effective for removing
the iron hydroxide lms from the pyrite surface in the range of pH
611 [90,142,143]. Since the iron hydroxide species are thermodynamically unstable in EDTA solutions of this pH range
[45,117,131], these species dissolve in EDTA solutions through
forming EDTA-iron complexes according to bellow reactions
[131]:
Fe(OH)3(s) + EDTA4 = FeEDTA + 3(OH)

(13)

Fe(OH)3(s) + EDTA4 = FeEDTA(OH)2 + 2(OH)

(14)

Fe(OH)2(s) + EDTA4 = FeEDTA2 + 2(OH)

(15)

In addition, the amount of the iron hydroxide species dissolved


in EDTA solutions increases with increasing Eh [45,144]. In the
acidic pH range, however, the different results are obtained in the
presence of EDTA. In this pH range, due to the adsorption of EDTA
and/or Fe-EDTA complexes on the pyrite surface, the dissolution
rate of the iron hydroxide species decreases dramatically. It has
also been noted that EDTA concentration should be optimum. The
presence of excess EDTA causes destroying sulfur-rich layers on the
pyrite surface after removing the iron hydroxide species which
results in decreasing the pyrite oatability [131]. To create the
optimum conditions and to completely eliminate the undesired
oxidation products, the use of non-oxidizing gas such as nitrogen
as a more effective method has been proposed by several
researchers [53,63,140].
It has been found that in industrial otation processes, pyrite
shows stronger oatability compared to experimental processes.
The studies indicated that vigorous pulp stirring and high solids
content in industrial otation cells cause generating intensive
particleparticle abrasion. This results in removing the hydrophilic
oxidation species from the pyrite surface and the exposure of the
underlying hydrophobic sulfur-rich layer, thereby increasing the
oatability of pyrite [90].
Pyrite activation
It has been obviously observed that pyrite can oat in the
presence of lead and copper minerals, even in alkaline solutions
[63,122,145,146]. This is due to the fact that pyrite can be activated
by lead or copper ions released from lead or copper containing
minerals [93,122,147]. The activation is one of the most important
phenomena associated with the otation of suldes. In this
process, metal species are precipitated or adsorbed on the mineral
surface, to create suitable sites for adsorption of the collector [85].
Pyrite activation by copper
Several studies have been conducted on adsorption mechanisms of copper ions on the pyrite surface. Some researchers have
proposed an ion exchange mechanism between ferrous ions and

EDTA: ethylene diamine tetra acetic acid.

Please cite this article in press as: H. Moslemi, M. Gharabaghi, A review on electrochemical behavior of pyrite in the froth otation process, J.
Ind. Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.12.012

G Model
JIEC 3218 No. of Pages 18

H. Moslemi, M. Gharabaghi / Journal of Industrial and Engineering Chemistry xxx (2016) xxxxxx

cupric ions in a system involving pyrite and copper ions (Reaction


(16)) [49,145,148,149].
FeS2 + Cu2+ = CuS + Fe2+ + S0

(16)

Although the Fe surface species have been known to be


energetically favorable for surface redox processes, it has been
reported that these species do not appear to have an electrochemical role in the Cu activation process of pyrite [150]. Chandra et al.
[150] studied Cu activated pyrite in slightly acidic conditions using
PEEM4 analysis and found that Cu is adsorbed onto the pyrite
surface in Cu+ form. It was also observed that greater amounts of
Sn2 and SOH species and smaller amounts of S2 and S22 are
present on activated pyrite relative to non-activated pyrite. It was
suggested that this may be due to the presence of O2/H2O or the
adsorption processes of cupric ions [150] which cause oxidation
during conditioning and activation of pyrite [150152]. Their
ndings conrmed the fact that the ion exchange is performed
along with the spontaneous reduction of cupric ions into cuprous
ions and the oxidation of sulfur [148150,153156]. The reduction
of cupric ions to cuprous ions has been known to be responsible for
a small increase in the pulp potential that is observed during the
activation of pyrite by copper [40,154]. It has been suggested that
covellite was the major compound on the surface of the Cu
activated pyrite [49,145,148,153,157159]. This was conrmed by
the results of XPS analysis [153] and voltammetry studies which
conducted on the activated pyrite [49]. It has also been reported
that copper is adsorbed only on the surface and does not penetrate
into the bulk of pyrite [45,146,160].
Studies conducted by Richardson et al. [161] showed that in
slightly acidic pH conditions, activation of pyrite by cupric ions is
strongly inuenced by the electrochemical potential of the pulp.
They observed that by increasing the electrochemical potential, the
adsorption of cupric ions decreases. This nding is in agreement
with studies by Wang et al. [158] who reported that O2 purging
causes a decrease in the available surface area for the copper
adsorption because of an increase in the oxidation of the pyrite
surface.
The activation of pyrite by copper in alkaline conditions has also
been investigated. It was found that in alkaline pH conditions, Cu2+
is rst adsorbed on the surface as copper hydroxide [162]. Then,
the copper hydroxide reacts with sulde (S22) at the pyrite surface
and is subsequently reduced to form Cu(I)S species through
oxidation of sulde to S21 or Sn2 (polysulde). It was also noted
that under alkaline conditions and for a strongly oxidized pyrite
surface, the rate of the copper adsorption is low because copper
has to penetrate into the surface hydroxide layer.
Weisener and Gerson [148] studied the activation of pyrite at
high electrochemical potentials using ToF-SIMS5 , EXAFS6 , XPS and
angle resolved XPS and reported that at alkaline pH values, the
pyrite surface, is initially rapidly activated by forming the Cu+sulde species but these species are then covered by precipitates of
cupric carbonate/hydroxyl. Although it has been suggested that the
cupric hydroxyl/carbonate species can form cuprous xanthate
upon the addition of xanthate collector to enhance the pyrite
otation [159,163], it is clear that these species are not able to react
signicantly with the xanthate within the time frame of the
otation process, and thus the otation recovery will decrease
[122].
The effects of the electrochemical potential on the pyrite
activation by copper ions at pH 9 and in the potential region from
100 to +400 mV (SHE) was studied and it was found that in

4
5
6

PEEM: photoemission electron microscopy.


ToF-SIMS: time of ight secondary ion mass spectrometry.
EXAFS: extended X-ray absorption ne structure.

alkaline conditions, the copper concentration on the surface was


independent of the electrochemical potential. They observed no
changes in the chemical state of sulfur on the pyrite surface by
potential changing [153].
The concentration of various copper species at different
electrochemical potentials at pH 9 was measured by Peng et al.
[122]. They found that the concentration of copper species strongly
depended on the electrochemical potential. It was observed that at
the potential of 185 mV (SHE) almost all aqueous copper (>99 wt.
%) was present as cuprous ion while at the potential of 10 mV
(SHE) this decreased to 28 wt.% and at the potential of +260 mV
(SHE), no aqueous Cu was present as cuprous ion. Moreover, it was
also found that if it was possible to form Cu(OH)2 precipitation,
almost all the copper could precipitate as Cu(OH)2 at potentials of
10 mV and +260 mV (SHE), while at the potential of 185 mV
(SHE), only a small amount of the copper could precipitate as Cu
(OH)2. According to these observations, it was concluded that
increasing pulp electrochemical potential caused an increase in the
formation of Cu(OH)2 and a decrease in the concentration of Cu+ on
the pyrite surface. On this basis, it has been suggested that since
the pyrite activation by copper ions strongly depends on the
formation of Cu+-sulde species, reducing conditions which
prevent oxidation of cuprous species to cupric species, are more
favorable for the activation process [161,164]. Peng et al. [122] also
noted that it can be possible to control the pyrite otation in the
presence of copper ions through the addition of reducing or
oxidizing agents.
Hicyilmaz et al. [49] performed contact angle measurements on
the copper activated pyrite in a potential range of 505 to +595 mV
(SHE) and at pHs of 4.67, 6.97 9.2 and 11. They observed that
activation by copper resulted in a signicant increase in the
hydrophobicity of the pyrite at all Eh and pH values. The pyrite
surface appeared to be similar to that of chalcopyrite when
activated with copper sulfate. It was reported that this was due to
the formation of CuS, Cu2S, and an iron decient surface layer
[118,136,148,153,159]. In this study, a slightly acidic pH that
resulted in the highest contact angles at all potentials was known
to be favorable for the otation of pyrite. The contact angles
decreased gradually with increasing pH and passing from slightly
acidic to neutral and alkaline conditions. The contact angle
measurements also showed that the pyrite surface was hydrophobic even at the potential of 505 mV (SHE). This could be attributed
to the presence of the elemental sulfur on the surface. Hicyilmaz
et al. [49] suggested that when copper ions interact with pyrite, the
elemental sulfur species are formed on the pyrite surface, and since
these species are reduced very slowly, the sulfur may still be
present on the surface in its elemental form at the potential of
505 mV (SHE). With increasing the potential to +385 mV (SHE), it
was observed that the contact angles rst decreased and then
increased. The decrease in the contact angles was known to be due
to the formation of HS and H2S and the increase in the contact
angles was explained by oxidation CuS into Cu2+ and S0. With a
further increase in the potential to +595 mV (SHE), a slight
decrease in the contact angles was also observed that could be
attributed to the formation of copper oxide/hydroxide species [49].
As mentioned above, under alkaline conditions and at high
electrochemical potentials, cupric species can depress the pyrite
otation by precipitation of the copper carbonate/hydroxyl.
Therefore, the overall adsorption of copper ions necessarily does
not mean that the activation of pyrite has effectively occurred
[122]. Based on XPS, IR7, UVvis8 and otation studies, Shen et al.
[165] proposed an electrochemical mechanism for the depression

7
8

IR: infrared spectroscopy.


UVvis: ultravioletvisible spectroscopy.

Please cite this article in press as: H. Moslemi, M. Gharabaghi, A review on electrochemical behavior of pyrite in the froth otation process, J.
Ind. Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.12.012

G Model
JIEC 3218 No. of Pages 18

H. Moslemi, M. Gharabaghi / Journal of Industrial and Engineering Chemistry xxx (2016) xxxxxx

of pyrite involving the formation of copper hydroxide on the


copper activated surfaces of pyrite by reduction of oxygen, which
leads to a decrease in the solution potential (Reactions (17) and
(18)).
H2O(aq) + 2e + 1/2O2(aq) = 2OH(aq)

(17)

Cu2+(aq) + 2OH(aq) = Cu(OH)2(surface)

(18)

Under such conditions, oxygen conditioning has been found to


have an important effect on the rate of the pyrite depression by Cu2
+
species [165].
Formation of the copper hydroxide on the surface may also
occur through galvanic interactions between the pyrite mineral
surface and copper sulde layer formed on the pyrite surface [165].
On this basis, cupric ions are produced by oxidation of the copper
sulde layer and then, react with hydroxide ions which are
produced by reduction of oxygen on the pyrite surface.
Pyrite activation by lead
The effect of the electrochemical potential on the activation of
pyrite by lead ions has also been investigated. Peng and Grano
[154] measured the pulp potential during the lead activation of
pyrite but observed no changes in the pulp potential. Peng et al.
[122] conducted zeta potential measurements on the leadactivated pyrite and found that the lead-activated pyrite displays
similar zeta-potential properties at different electrochemical
potentials. They also found that the lead-activated pyrite has an
iso-electric point similar to lead hydroxides, oxides or carbonates
[166,167]. On the other hand, it was observed that almost all the
lead ions added during the activation process were extractable by
EDTA solutions. These ndings obviously show that the activation
of pyrite by lead occurs mainly through the formation of lead
surface complexes such as hydroxides and, unlike the activation
process by copper, new metal sulde phase does not form [154].
Therefore, since the activation of pyrite by lead ions does not
appear to occur through an electrochemical mechanism, in order
to control pyrite otation in the presence of lead ions, other
methods apart from the methods based on the electrochemical
processes have to be sought [122].
Collector adsorption on pyrite surface
Due to semiconducting properties of sulde minerals, adsorption of collectors on their surface is usually electrochemical in
nature and affected by the electrochemical conditions of the
otation pulp [39,92,99,168170]. Collector adsorption is inuenced not only by oxygen, but also by all the oxidizing and reducing
agents present in the pulp and collector adsorption mechanism
depends on activity of the mineral surface [171,172].
Collector adsorption on non-activated surface
The general proposed mechanism for hydrophobisation of the
pyrite surface in the presence of xanthate collectors is the
oxidation of xanthates to hydrophobic dixanthogens and their
adsorption on the surface [173178] according to Reaction (19)
[179,180].
2X = X2 + 2e

(19)

Charge balance for the oxidation reaction of xanthate to


dixantogen is preserved by the cathodic reduction of adsorbed
oxygen according to Reaction (17) [181183].
Although dixanthogens have been known as the main
hydrophobic products formed on the pyrite surface in the presence

of xanthates [85], the results of otation studies conducted by


Trahar [184] using ethyl xanthate as collector showed that pyrite
can be oated by xanthates at potentials much more cathodic than
the equilibrium potentials of the xanthate/dixanthogen couples.
Thus, dixanthogens cannot be solely responsible for the otation of
pyrite. Using the Eh-pH stability diagram of the FeEXH2O
system, it was clearly found that the potential of otation
commencement is approximately equal to potential for formation
of ferric xanthate compounds [181,185187]. Therefore, it can be
concluded that the formation of the ferric xanthate compounds
and adsorption of ethyl xanthate ions may be the reason why the
pyrite otation occurs at more cathodic potentials than the
potential required for the dixanthogens formation from xanthates
[185,188,189]. FTIR9 studies showed that the adsorption of both
ferric xanthate compounds and dixanthogens occurs on the pyrite
surfaces reacted with xanthates [85,99,146,150,159,185,186,190,
191], but the amount of adsorbed ferric xanthates is much less than
that of adsorbed dixanthogens [85,99,159,169,185,192,193]. It was
also found that ferric xanthates form a monolayer on the surface,
whereas dixanthogens are adsorbed as multiple layers [186,194]. In
addition to ferric xanthates and dixanthogens, a minor amount of
monothiocarbonates was also found on the surface by voltammetry studies conducted by Harris and Finkelstein [195]. They
identied them as intermediate products of the ferric xanthates
formation during interaction between pyrite and xanthates.
Using FTIR studies, four main steps in the adsorption process of
xanthate collectors on the pyrite surface were identied. These
steps include [186]:
1 - Surface oxidation of pyrite.
2 - Xanthate ions adsorption on pyrite surface and formation of
ferric xanthates.
3 - Xanthate ions oxidation and dixanthogen formation.
4 - Dixanthogen adsorption on pyrite through ferric xanthates
present on the surface.
Based on the literatures, it was found that pure dixanthogen
cannot be effectively adsorbed on pyrite without the existence of
suitable adsorption sites on the surface [177,186]. The ferric
xanthate compounds are believed to be one of the suitable sites for
the adsorption. Thus, their formation on the pyrite surface appears
to be necessary for the effective adsorption of dixanthogens [186].
Xanthates are reducing agents [11,177,181]. Due to the
equilibrium potential of the xanthate/dixanthogen couple, as
more cathodic than the mixed potential of pyrite, the xanthate
addition causes a decrease in the mixed potential of the pulp
[40,181]. The magnitude of this decrease in the potential indicates
that the intensity of interaction of the xanthate with pyrite [40].
Dixanthogen is formed on the mineral surface only when the
mixed potential of the pulp rises above the xanthate/dixanthogen
equilibrium potential [128,181,196]. Pyrite is the most noble among
the sulde minerals. Therefore, more dixanthogen is expected to
be formed on the pyrite surface compared to other sulde minerals
[99]. The oxidation potential of dixanthogen (EhX2) can be
calculated by the Nernst equation (Relation (20)) [85].
EhX2 = E0X/X2  0.059log[X]

(20)

E0X/X2

is the standard potential of the half-cell reaction


Where
expressed by Equation (19) and [X] is the molar concentration of
xanthate. It was demonstrated that there is an inverse relationship
between the equilibrium potential of the xanthate/dixanthogen
redox reaction and length of the hydrocarbon chain. An increase in
the length of the hydrocarbon chain causes a decrease in the

FTIR: fourier transform infrared spectroscopy.

Please cite this article in press as: H. Moslemi, M. Gharabaghi, A review on electrochemical behavior of pyrite in the froth otation process, J.
Ind. Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.12.012

G Model
JIEC 3218 No. of Pages 18

H. Moslemi, M. Gharabaghi / Journal of Industrial and Engineering Chemistry xxx (2016) xxxxxx

equilibrium potential of the xanthate/dixanthogen couple


[185,186].
The rate of the xanthate adsorption by pyrite from aqueous
solutions can be expressed as Relation (21) [174,187,197].
(dcX/dt) = kcX-a  cO2bcH+g

(21)

1/2

where k = (1/V)k'(AaAc) ,V is the liquid volume, k' is the rate


constant, Aa and Ac are the surface areas of sites over which the
oxidation and reduction reactions take place, respectively, cX-, cO2
and cH+ are the xanthate ion, dissolved oxygen and hydrogen ion
concentrations in the solution, respectively, and a, b, and g are
reaction orders. For example, on pyrite at a constant oxygen
concentration, a = 0.5 and g = 0 in neutral solution and a = 0.75 and
g = 0.25 in alkaline solution [187].
Apart from xanthates, other collectors such as DTPI10 are also
used in the sulde minerals otation [62,198200]. Flotation
studies in the presence of these collectors have shown that the
electrochemical conditions have signicant inuence on their
performance [198,199]. Gler [47] and Hicyilmaz et al. [48],
performed spectroscopic and electrochemical studies and found
that DTPI dimer can only be formed on the pyrite surface at highly
oxidizing conditions. Despite these studies, the electrochemical
behavior of pyrite in the presence of these types of collectors is still
not well understood [62].
In another study, researchers have conducted otation tests on
pyrite using two different types of collectors, including PAX and a
trithiocarbonate. The results of the otation experiments showed
that lower recoveries were obtained by trithiocarbonate collector
type compared to PAX [201]. These differences in the otation
results could be due to the differences in the electrochemical
properties of these two types of collectors. It was found that the
standard reduction potentials of xanthate type collectors are about
90 mV more negative than that of their corresponding trithiocarbonate homologues. On this basis, it could be concluded that during
otation, pulp electrochemical potential was high enough for
xanthate oxidation to dixanthogen and its adsorption on the pyrite
surface, whereas this value of potential was probably not enough for
the trithiocarbonate type collector oxidation. Therefore, adsorption
of trithiocarbonate oxidation products did not occur on the pyrite
surface, leading to the recovery reduction. These results clearly
indicated that it is necessary to consider the electrochemical nature
of collectors used in the otation processes of pyrite.
Effect of pH on collector adsorption
The effect of pH on the xanthate oxidation and formation of
dixanthogen has also been studied. The results showed that, at
lower pH levels, pyrite surface is oxidized at more positive
potentials than the equilibrium potential for the dixanthogen
formation. Under such conditions, the oxygen reduction and the
xanthate oxidation are the dominant electrochemical reactions in
the system and as a result, the pyrite surface will be hydrophobic
through adsorption of dixanthogen. By increasing pH, the required
potential for oxidation of pyrite decreases. At higher pH levels, the
oxidation of the pyrite surface occurs at lower potentials than the
potential required for the dixanthogen formation from xanthate.
Therefore, the major electrochemical reactions in the system
include the oxidation of pyrite and the reduction of oxygen. In such
a state, rate of the xanthate oxidation decreases and the pyrite
surface will remain hydrophilic [50].
It has been suggested that the negative effect of pH increasing
on the xanthate adsorption can be reduced by increasing xanthate
concentration. The results of voltammetry studies conducted by
Janetski et al. [50] revealed that the oxidation potentials of pyrite

and xanthate are cathodically shifted by an increase in the pulp pH


and xanthate concentration, respectively. However, the shift
caused by increasing xanthate concentration by an order of
magnitude is greater than that by increasing pH by a unit. These
results showed that it is possible to nd a relationship between the
pulp pH and the concentration of a xanthate collector during a
otation process. For example, they represented the general
Relation (22) for the ethyl xanthate collector.
[C2H5OCS2]/[OH]0/8 = Constant

PAX: potassium amyl xanthate.

(22)

Effect of oxygen content on collector adsorption


Some researchers have studied the effect of oxygen on the
xanthate adsorption on the pyrite surface. It was found that the
presence of oxygen improves the xanthate adsorption by pyrite
[174,202,203]. An increase in the dissolved oxygen concentration
in the pulp causes an increase in the pulp potential and
consequently an increase in the oxidation rate of xanthate to
dixanthogen on the pyrite surface [174,202,204,205]. Furthermore,
it was also suggested that oxysuldes and also ferric ions produced
by the pyrite oxidation can facilitate dixanthogen formation
[53,85,190,206,207]. Investigations showed that the xanthate
adsorption on the non-activated pyrite surface was along with
an increase in ferrous iron concentration in the pulp [208]. It was
found that the reduction of ferric hydroxide to ferrous ions and
development of dixanthogen occur simultaneously on the pyrite
surface [146,208]. This nding is in agreement with results of FTIR
studies conducted by Leppinen et al. [135]. Fig. 3 shows this
process schematically.
Although it is believed that oxygen is necessary for the pyrite
otation in the presence of xanthate, the ndings of some studies
have shown that the recovery of pyrite may be improved when
nitrogen is used instead of oxygen as otation gas [53,209,210].
Flotation experiments in the presence of 50 g/t PAX11 collector by
air and nitrogen during the conditioning step and as otation gas
showed that the pyrite recovery increased from 31.4% to 94.1%
when air was replaced by nitrogen [53]. Using DRIFT12 analysis,
they found that the xanthate adsorption and the dixanthogen
formation increased in the nitrogen atmosphere. Pulp potential
measurements showed that this increase in the formation of
dixanthogen was due to lowering the pulp potential and
consequently inhibiting the formation of the hydrophilic oxidation
products on the pyrite surface because of the presence of nitrogen.
Similar results were also obtained in experiments performed by
Miller et al. [211]. On the basis of XPS analysis, they reported that
the nitrogen atmosphere not only inhibited the formation of
hydrophilic oxidation products on the pyrite surface, but also
caused reduction of the surface compounds on pyrite and as a
result, the xanthate adsorption on pyrite surface was facilitated.
Effect of pyrite surface oxidation products on collector adsorption
Electrochemical polarization method was used to determine
the effect of the pyrite surface composition on the xanthate
adsorption in a otation system. The results of otation on the
electrochemically polarized pyrite samples (at polarization potentials of +250 and 250 mV (SHE)) in the alkaline environment (pH
9.18) showed that the anodic polarization decreased pyrite
recovery by 12%, whereas the cathodic treatment increased the
recovery by 45% [212]. Using laser scanning microscopy, it was
found that the adsorption of xanthate on the anodically polarized
sample signicantly decreased due to severe oxidation of surface
and its coverage by iron oxidation products. In contrast, in the case

11
10

12

DRIFT: diffuse reectance infrared fourier transform spectroscopy.


DTPI: dithiophosphinate.

Please cite this article in press as: H. Moslemi, M. Gharabaghi, A review on electrochemical behavior of pyrite in the froth otation process, J.
Ind. Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.12.012

G Model
JIEC 3218 No. of Pages 18

10

H. Moslemi, M. Gharabaghi / Journal of Industrial and Engineering Chemistry xxx (2016) xxxxxx

Fig. 3. Dissolution of surface ferric hydroxide and dixanthogen formation on non-activated pyrite surface [208].

of the cathodically treated sample, analysis of the microimages


showed that not only was no new phase due to the iron oxidation
formed on the surface, but the iron oxidation products present on it
were also dissolved. Instead, it was observed that the formation of
dixanthogen phases on the pyrite surface increased after contacting the sample with xanthate. This nding that the electrochemical
adsorption of xanthate on the pyrite surface occurs effectively only
in slightly oxidizing or reducing environments can be attributed to
facilitation of electron transfer across the interface between the
slightly oxidized or the unoxidized surface and solution. The
results of these studies and similar results clearly indicate the
importance of the pyrite oxidation products in the adsorption of
xanthate on the surface and its oatability.
Effect of pyrite type and origin on the collector adsorption
It has been found that a defect in the pyrite lattice inuences the
adsorption process of xanthate on the pyrite surface. Electronic
structure and band structure studies on four types of pyrite including
perfect, Co-substituted, As-substituted, and intercrystalline Au
pyrites was performed by Xian et al. [106] and it was found that
there is a relationship between the amount of xanthate adsorbed on
the pyrite surface and band-gap value. The results of the studies
showed that a pyrite with a lower band-gap transfers an electron
from a xanthate collector to an oxidant more quickly (such as oxygen
or ferric iron). As a result, the oxidation of xanthate collector to
dixanthogen occurs more readily. Moreover, they found that the
stability of the electrochemically adsorbed oxidation products on the
surface increases with a decrease in the band-gap of pyrite [116,213].
It was also observed that intercrystalline Au and co-substituted
pyrites adsorbed more xanthate compared to perfect and Assubstituted pyrites due to having lower band-gap values [106].
Xanthate collectors are also used to separate coal-pyrite in the
coal desulfurization process by otation. Adsorption experiments
and rest potential measurements conducted by Miller et al. [214]
showed that coal-pyrite has more adsorption ability for xanthate in
comparison to pure pyrite. It appears that the rest potential of coalpyrite is inuenced by other electrochemical processes such as
oxidation of electrochemically active components present in coal.
Collector adsorption on activated surface
Collector adsorption on copper activated surface
During conditioning of pyrite, it was observed that a yellow
precipitate was formed on the pyrite surface immediately after

the addition of KEX13 collector at the end of the period of the


activation with copper ions [150]. It was also observed that the
amount of this precipitate increased by increasing collector
dosage. Moreover, the yellow precipitate was found not to be
removed, even by washing with a sulfuric acid solution (pH 5). FTIR
studies revealed that this precipitate was comprised mainly of
cuprous xanthate along with a small proportion of dixanthogen.
The formation of cuprous xanthate on the copper-activated pyrite
surfaces in the presence of xanthate has been reported by other
researchers [45,146,153,159,162].
The proposed mechanism for the xanthate adsorption is a twostep process, including a surface electrochemical reaction which
produces copper ions and a chemical reaction that produces the
cuprous xanthate compound [85]. The cuprous xanthate compound is formed as a monolayer on the pyrite surface and known as
the main hydrophobic product of the interaction of xanthate with
the copper-activated surfaces [45,146,153,159,162].
It was reported that, during conditioning of pyrite at pH 7, if
equal dosage of xanthate collector and copper are used, the
cuprous xanthate will be the only formed compound. However, if
unequal amounts of xanthate and copper are used, considerable
amounts of dixanthogen will also be formed in addition to the
cuprous xanthate [146]. The studies have shown that, at higher
concentrations of copper, cupric xanthate is also formed from
combining aqueous Cu2+ ions with xanthate, but due to its unstable
nature, it is then oxidized to dixanthogen [150,163].
Xanthates show a strong afnity to copper and their adsorption
on the copper- activated pyrite surfaces increases with increasing
Eh. However, it has been found that if Eh highly increases, the
extensive pyrite oxidation will occur and the surface will be
covered with iron hydroxide/oxide lms. As a result, due to the
hydrophilic nature of these products, the pyrite oatability will
decrease [45].
Collector adsorption on lead activated surface
In the case of lead activated pyrite surfaces, it was found that
lead has a catalytic effect on the dixanthogen formation. The
proposed mechanism for the dixanthogen precipitation includes
three steps. First, xanthate adsorbs on the activated sites as
Pb(OH)X. Then, the xanthate ions adsorption occurs to form

13

KEX: potassium ethyl xanthate.

Please cite this article in press as: H. Moslemi, M. Gharabaghi, A review on electrochemical behavior of pyrite in the froth otation process, J.
Ind. Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.12.012

G Model
JIEC 3218 No. of Pages 18

H. Moslemi, M. Gharabaghi / Journal of Industrial and Engineering Chemistry xxx (2016) xxxxxx

Pb(OH)X.X. Finally, the adsorbed xanthate is oxidized and


dixanthogen is precipitated [85].
Depressants
Inorganic depressants
Cyanide, mainly as potassium cyanide (KCN) and sodium
cyanide (NaCN) has been known as one of the most common
depressants for pyrite in otation process [215]. Simple cyanide
compounds are dissolved in water and produce ionic species
according to the following reaction [216]:
A(CN)x = A

+x

+ x(CN)

(23)

where A is an alkali metal and X is the valence of the alkali metal


and it is equal to the number of cyano groups present in the
compound.
The released cyanide anion, CN, hydrolyzes through abstracting a proton from water to produce aqueous hydrogen cyanide
[217]:
CN + H2O = HCN + OH

(24)

Hydrogen cyanide is a volatile substance and can easily


evaporate from the solution as a very toxic gas [218220]. The
concentrations of cyanide ion and hydrogen cyanide are governed
by pH of the solution. These concentrations are at equilibrium
when the pH is equal to pKa of HCN (acidbase dissociation
constant) [216]. At pH values higher than the pKa value, cyanide
ion concentration in solution increases while at pH values lower
than the pKa value, volatile hydrogen cyanide concentration
increases [220]. Since pKa value depends on various factors such as
the concentration of dissolved solids in the solution, reported
values for pKa is often inconsistent [216]. Although a value equal to
9.31 has been frequently quoted as pKa in several reports [221
224]. Therefore, it is necessary to keep the solution pH above
9.31 to avoid volatilization of cyanide as hydrogen cyanide gas
[225].
It has been found that electrochemical interaction of xanthates
is strongly affected by the presence of cyanide ions [173]. Results of
voltammetry studies conducted by Janetski et al. [50] revealed that
at constant xanthate concentration and pH, increasing cyanide ion
concentration causes a shift in the oxidation potential of xanthate
to more positive values. Based on these results, they reported that
cyanide ions appear to have an inhibiting effect on the oxidation
process of xanthate. It was also found that, at constant xanthate
concentrations, the anodic shift in the oxidation potential of
xanthate due to increasing the cyanide ion concentration is
gradually reduced with decreasing the solution pH. These ndings
have been conrmed by other studies [181,194].
De Wet et al. [181] by electrochemical impedance measurements, observed that in a low frequency range, the electrochemical
impedance of the pyrite electrode in the presence of NaCN was
considerably higher than in its absence. This shows that the
presence of cyanide increases polarization resistance. An increase
in polarization resistance means a decrease in the electrochemical
reaction rates on the pyrite surface. De wet et al. [173] observed
that the mixed potential of pyrite decreased from +195 to +125 mV
(SHE) after adding 200 mg/l cyanide to the otation pulp in the
presence of 50 g/ton of xanthate. Since by decreasing mixed
potential, the oxidation rate of xanthate to dixanthogen decreases,
it was concluded that suppressing the xanthate oxidation in the
presence of cyanide can be due to decreasing the mixed potential
[173,181,188].
There are two possibilities that can cause a decrease in the
mixed potential in the presence of cyanide. One possibility is to

11

inhibit the surface reduction reactions [181]. However, no evidence


has been found that cyanide inhibits the reduction reaction of
oxygen [173]. Another possibility is to enhance oxidation reactions
[181]. Since cyanide has a strong reducing nature, it is oxidized to
cyanate and/or cyanogen according to relations (25)(27) [188].
2CN + O2 = 2CNO

(25)

2CN + H2O + 1/2O2 = (CN)2(g) + 2OH

(26)

(CN)2 + 1/2O2 + 2OH = 2CNO + H2O

(27)

However, it has been found that these reactions are slow, and
thus do not appear to have an important role in suppressing the
xanthate oxidation through lowering the mixed potential [181].
Enhancing the anodic reactions by cyanide may also be due to
the formation of soluble iron cyanide compounds instead of the
insoluble oxide/hydroxide species. Since these iron cyanide
compounds, unlike the oxide/hydroxide species, are soluble, they
cannot inhibit the pyrite oxidation, and thus the anodic reactions
are enhanced [181,188]. This nding was conrmed by voltammetry studies conducted by Wang and Forssberg which showed a
linear relationship between the concentration of cyanide and the
peak current of the pyrite oxidation [188]. Although these studies
showed that cyanide suppresses xanthate oxidation through
decreasing the mixed potential, it was found that the xanthate
oxidation in the presence of cyanide is not enhanced by increasing
mixed potential by adding an oxidizing agent [173]. Therefore, the
decrease in the mixed potential does not appear to be the main
cause for suppressing the xanthate oxidation.
Several other causes such as high consumption of oxygen in the
cyanide oxidation reactions and direct reduction of dixanthogen to
xanthate by cyanide have also been reported [188]. The most
acceptable explanation for the inhibition of the xanthate oxidation
by cyanide has been found to be the formation of insoluble
cyanoferrate complexes on the pyrite surface [50,99]. It has been
found that iron hexacyanide compounds can be formed by the
reduction of the surface ferric hydroxide layer in the presence of
cyanide [50] according to Reaction (28) [99].
5Fe(OH)3 + 12CN + 15H+ + 3e = Fe3[Fe(CN)6]2 + 15H2O

(28)

It has also been reported that, in the presence of cyanide,


insoluble ferric xanthate compounds formed on the pyrite surface,
which act as suitable adsorption sites for dixanthogen, become
unstable. These compounds are then replaced by Fe(CN)64 and
also Fe(CN)63 at higher electrochemical potentials (Reaction (29))
[188].
Fe(OH)2EX(s) + 6CN = Fe(CN)63 + EX + 2OH

(29)

Consequently, the dixanthogen adsorption on the surface is


reduced. The presence of these species on the surface has been
conrmed by zeta and redox potential measurements and also
thermodynamic calculations [226]. In addition, it has also been
found that since these species are hydrophilic, their formation on
the surface results in a decrease in the hydrophobicity due to the
adsorbed dixanthogen [50]. It has been noted that cyanide
effectively acts as a depressant only when the pyrite surface
undergoes extensive oxidation [50,227].
Studies have shown that, even in the presence of copper ions,
cyanide is able to depress pyrite. It has been observed that, in
alkaline conditions and in a potential range of approximately
300 to +450 mV (SHE), addition of cyanide to the otation pulp
causes the formation of cyano-copper species such as insoluble
CuCN and Cu(CN)32 at low and high CN/Cu concentration ratios,

Please cite this article in press as: H. Moslemi, M. Gharabaghi, A review on electrochemical behavior of pyrite in the froth otation process, J.
Ind. Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.12.012

G Model
JIEC 3218 No. of Pages 18

12

H. Moslemi, M. Gharabaghi / Journal of Industrial and Engineering Chemistry xxx (2016) xxxxxx

Table 2
Mechanisms of depression of pyrite otation by cyanide ions in different conditions.
Condition

Mechanism

Reference

Collector-less otation
In the presence of xanthate
In the presence of copper ions

Removing hydrophobic sulfur species from the surface by the formation of thiocyanate compounds
Replacing surface adsorption sites for dixanthogen by insoluble cyanoferrate complexes
Reducing concentration of the free copper ions in the pulp and inhibiting the formation of
coprous xanthate by the formation of cyano-copper species
Cyanide is not able to depress the pyrite otation.

[188]
[50,99]
[188]

In the presence of lead ions

respectively. The formation of these species leads to a signicant


reduction in the free copper ions concentration in the pulp and
consequently, inhibition of the coprous xanthate formation. It has
also found that, even if the coprous xanthate compounds are
formed on the surface, the addition of cyanide will dramatically
reduce the stability of these compounds and will dissolve them
(Reactions (30) and (31)) [188].
CuEX(s) + CN = CuCN(s) + EX

(30)

CuEx(s) + 3CN = Cu(CN)32 + EX

(31)

It has been reported that cyanide, even at relatively high


concentrations, is not able to depress the lead activated pyrite
because the cyanide anion cannot form complex with the lead
cations [228]. This nding has been used to improve performance
of N2TEC otation technology applied at Newmont's Lone Tree
Plant in Nevada, where it was observed a loss in the auriferous
pyrite recovery, due to the presence of cyanide in the otation mill
water. It has been found that higher recovery can be obtained at
lower xanthate concentrations in reducing conditions by the
activation with lead ions [229].
In collectorless otation, the elemental sulfur and polysuldes
have been found to be responsible for the pyrite oatability. In the
presence of cyanide, these species are unstable [188]. It has been
found that cyanide is able to react with the elemental sulfur and
polysuldes to form thiocyanate and to remove the hydrophobic
species from the surface in the range of low potentials (Reactions
(32) and (33)).
S + CN = SCN

(32)

(x  1)CN + Sx2 = (x  1)SCN + S2

(33)

These experimental ndings have been conrmed by thermodynamic calculations [188] and also by XPS analysis [68]. Table 2
summarizes the mechanisms of depression of the pyrite otation
by cyanide ions in different conditions.
Sulfoxy reagents such as sulfur dioxide, sodium sulte or
sodium metabisulte are also widely used as depressants for pyrite
[50,146,202]. Studies have shown that these reagents suppress the
adsorption of xanthate by reducing the mixed potential to below
the potential required for the oxidation of xanthate to dixanthogen
[50,202,230] and enhancing formation of hydroxide species on the
pyrite surface [146].
Organic depressants
Although various inorganic reagents (e.g., cyanide and sulfoxy
species) have been found to be effective in the depression of pyrite,
there are problems associated with the use of these reagents such
as high cost and environmental concerns [3,91,231]. In order to
overcome such problems, some researchers have focused on
developing alternative cheap and environmental friendly organic
depressant.
Natural polysaccharides, such as starch, dextrin and guar gum,
are used as depressant. These organic compounds are polymers

[228,229]

with high molecular weights composed of monosaccharide units.


Natural polysaccharides depress the pyrite in the otation through
selective adsorption on pyrite surface [232]. It was found that both
natural polysaccharides and xanthats adsorb on the pyrite surface,
but hydrophobic molecules of dixanthogen are enveloped by large
hydrophilic molecules of natural polysaccharides, resulting in the
pyrite depression [232234]. Due to the heterogeneous nature of
the oxidized pyrite surface, their adsorption occurs at different
sites on the surface. Xanthate ions adsorb on non-oxidized sites on
the pyrite surface, while these organic depressants adsorb on
oxidized sites [233]. The adsorption of natural polysaccharides on
oxidized sites was found to be due to the existence of two hydroxyl
groups on C-2 and C-3 carbon atoms in each unit of the
polysaccharides, which enable polysaccharides to form complexes
with hydroxide species present on the surface. Therefore,
adsorption of these organic compounds on pyrite surface resulting
in pyrite depression depends on level of surface oxidation and
density of metal oxidation species on the surface [232].
CMC14 is another organic compound which acts as depressant
in pyrite otation process. It is a cellulose derivative being
produced by etherication process. The presence of negatively
charged carboxyl groups in CMC structure in addition to hydroxyl
groups, causes increasing CMCs selectivity compared to natural
polysaccharides. Unlike the carboxyl groups, which is able to
interact with various forms of metallic species, the hydroxyl groups
is only able to interact with metal hydroxyl species. Therefore,
similar to natural polysaccharides, adsorption of this depressant on
the pyrite surface is also affected by intensity of the surface
oxidation and presence of oxidation products on the surface [232].
PAM15 (a synthetic polymer with general chemical formula
CH2CHCONH2) is another organic compound which, similar to
natural polysaccharides and CMC. PAM adsorption on the surface
depends on the presence of metal oxyhydroxy species which are
produced from oxidation process. Studies conducted in laboratory
showed ability of PAM and its derivatives to depress pyrite in
otation process [232].
Lignosulfonates, or sulfonated lignins (water-soluble anionic
polyelectrolyte polymers) are also used as organic depressants for
pyrite otation. Depression mechanism of pyrite otation by these
organic compounds has been investigated in both non-activated
and Cu-activated states [232]. Electrochemical studies conducted
on non-activated pyrite otation showed that the biopolymers,
after adsorption on pyrite, passivate surface and suppress
electrochemical reactions occurring on the pyrite surface, including oxidation and reduction reactions of pyrite itself and also
oxidation of xanthate on the surface [232,235237]. It was found
that this was due to form insulator layers on the surface, resulted in
a reduction in pyrite conductivity and an increase in the difculty
in electrons transferring between pyrite-solution and pyritexanthate interfaces [232,236,237]. In the otation of Cu-activated
pyrite, studies showed that the biopolymers was able to accelerate
cuprous sulde and xanthate compounds to form complexes with
extracted cuprous ions from the sulde and xanthate forms

14
15

CMC: carboxymethyl cellulose.


PAM: polyacrylamide.

Please cite this article in press as: H. Moslemi, M. Gharabaghi, A review on electrochemical behavior of pyrite in the froth otation process, J.
Ind. Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.12.012

G Model
JIEC 3218 No. of Pages 18

H. Moslemi, M. Gharabaghi / Journal of Industrial and Engineering Chemistry xxx (2016) xxxxxx

[232,238]. In these conditions, the pyrite depression occurs as a


result of the formation of Cu-biopolymer and Cu(OH)2 hydrophilic
species on the pyrite surface [232].
DETA16 is another organic compound which is used as a pyrite
depressant in the otation process. In the case of Cu-activate pyrite
otation, it was found that DETA causes depression of the pyrite
surface through promoting oxidation of Cu(I) to Cu(II) and
removing Cu(I)-S species from the surface resulting in a substantial
reduction in xanthate adsorption. DETA also increases hydrophilic
oxidation products such as Fe-Ox, Cu(OH)2 and Cu-DETA. Depression mechanism of Pb-activate pyrite by DETA has also been
investigated. Electrochemical studies indicated that DETA suppresses xanthate oxidation, probably by reducing rest potential of
Pb-activated pyrite [232].
Ovalbumin (chicken egg albumin), which is a high molecular
weight monomeric glycoprotein was proposed by Guler et al. [91]
as another organic reagent. Ovalbumin, due to having four
sulfhydryl (
SH) bonds in cysteine groups and one disulde
(S
S) bond in cysteine, has been found to be electrochemically
active in aqueous solutions [239242].
Flotation studies at pH 9.2 and in a potential range of 400 to
+800 mV (SHE) by ovalbumin as depressant showed that in
reducing conditions, adsorption of ovalbumin on the pyrite surface
occurred at lower rates. Since, in the alkaline conditions and at
potentials lower than the pyrite open circuit potential, both pyrite
and ovalbumin have negative surface charges, it was concluded
that the lower rates of the adsorption are due to the electrostatic
repulsion between them. In slightly oxidizing conditions, a slight
increase was observed in the adsorption rate, which was found to
be due to the hydrophobic interaction of ovalbumin with the
hydrophobic oxidation products formed on pyrite. Results showed
that the depressing effect of ovalbumin reaches its maximum at
oxidizing potentials around +500 mV (SHE) [91], where the surface
is intensely oxidized and the ferric species [58,63,99,243], being
favorable for the electrostatic interaction of the surface with
ovalbumin molecules [3,240,244], are formed on the pyrite surface.
The adsorption rate has been found to remain approximately
constant in the highly oxidizing environment. Results of cyclic
voltammetry and FTIR spectroscopy showed that in these
conditions, the pyrite depression occurs by the formation of
ovalbuminmetal chelates mainly by the electrochemical processes [91].
Effect of grinding environment on pyrite electrochemical
properties
Effect of grinding media type
It has been found that grinding has signicant effects on the
electrochemical potential at the mineral/solution interface and the
nature of reactions occurring in a mill is considerably inuenced by
the type of grinding media [122]. The use of electrochemically
active grinding media causes galvanic interaction between pyrite
and the grinding media during grinding, which results in an effect
on the electrochemical behavior of pyrite. In this galvanic
interaction, pyrite due to having a higher rest potential acts as a
cathode and results in the oxygen reduction on its surface and the
production of hydroxide ions. In contrast, the grinding media,
which act as an anode, are oxidized and release ferrous ions. The
produced ferrous ions undergo further oxidation to ferric ions and
then react with the hydroxide ions and are precipitated as ferric
hydroxide on the pyrite surface [72,245]. The iron oxidation
species emanating from the grinding media have been found to

16

DETA: diethylenetriamine.

13

have an important role in depressing the pyrite otation [245257].


It has also been reported that the use of the electrochemically active
grinding media causes a decrease in the xanthate oxidation rate in
the otation process [185,201]. This is due to the dissolved oxygen
consumption in pulp during the grinding media oxidation, which
results in a reduction in Eh of the pulp [146,201,258].
The extent of the galvanic interaction occurring during grinding
and consequently the magnitude of change in the electrochemical
potential of pulp has been found to be directly related to the
difference in the electrochemical properties between grinding
media and sulde minerals [72,128,245]. It was reported that a
conventional mill produces a reducing grinding environment with
a low electrochemical potential while a porcelain or an autogenous
mill produces a non-reducing grinding environment with a high
electrochemical potential [252]. Moreover, it was found that the
use of grinding media with more electrochemical activity results in
more reducing grinding conditions [93,147]. Chen et al. [259]
conducted otation tests on two samples containing pyrite that
were ground with two types of grinding media, including mild
steel and stainless steel media. The results showed that the
otation recovery of pyrite for the sample ground with less active
grinding media (stainless steel media) was signicantly higher
than that with more active grinding media (mild steel media).
Based on the XPS analysis, they found that the amount of oxidation
species on surfaces of the pyrite particles increased after grinding,
and this increase was signicantly greater in the case of the sample
ground with mild steel media. These results were in good
agreement with the results of EDTA extraction tests reported by
Peng et al. [93].
Studies have shown that different grinding media also have
different effects on the activation process of pyrite. It has been
found that grinding with more active grinding media results in
more pyrite activation by copper ions [259263]. In contrast, it has
been observed that electrochemically inert grinding media, which
produce a strong oxidizing grinding condition, have an enhancing
effect on the activation of pyrite by lead ions [122]. This difference
in the activation by copper and lead ions is attributed to the
different mechanisms of the pyrite activation by these ions. The
activation of pyrite by copper ions occurs through an electrochemical mechanism, which is enhanced in reducing conditions,
while the activation of pyrite by lead ions is independent of the
electrochemical potential [154]. Since in an operating mineral
processing plant, it is impractical to adjust the electrochemical
condition during grinding by replacing the existing mill and
grinding media, the addition of oxidizing or reducing agents has
been proposed by researchers as an alternative solution for
controlling the activation process of pyrite during grinding [122].
The effect of the electrochemical activity of media grinding on the
various processes in pyrite otation is summarized in Table 3.
Effect of oxygen content in grinding system
It is well known that oxygen plays an important role in the
galvanic interaction during grinding. Huang and Grano [261]
studied the oatation recovery of pyrite as a function of galvanic
current during grinding in the presence of copper ions under
different atmospheres including nitrogen, air and oxygen. Results
showed that nitrogen purging produced the lowest galvanic
currents as expected. In this condition, with increasing the galvanic
current, the pyrite recovery increased which appeared to be
associated with the increased rate of the cupric ion reduction on
the pyrite surface. When air was purged during grinding, higher
galvanic currents were recorded. This was mainly due to an
increase in the reduction rates of ferric ion and oxygen. An increase
in the pyrite recovery was still observed with the galvanic current
and the copper ions addition. This increase was found to be

Please cite this article in press as: H. Moslemi, M. Gharabaghi, A review on electrochemical behavior of pyrite in the froth otation process, J.
Ind. Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.12.012

G Model
JIEC 3218 No. of Pages 18

14

H. Moslemi, M. Gharabaghi / Journal of Industrial and Engineering Chemistry xxx (2016) xxxxxx
Table 3
Effect of electrochemical activity of media grinding on various processes in pyrite otation.
Process

Effect of increasing electrochemical activity of media grinding

Reference

Collector-less otation
Xanthate adsorption
Activation by copper
Activation by lead

A decrease in pyrite oatability


A decrease in xanthate adsorption rate
An increase in activation rate
A decrease in activation rate

[259]
[185,201]
[259263]
[122]

because of the increased pulp potential due to air purging,


resulting in an increase in the adsorption of xanthate in the
conditioning stage. In the oxygen purging, the galvanic currents
were the highest because of very high rates of the reduction of
ferric ion and oxygen. In this case, it was found that the amount of
the iron oxidation species on the pyrite surface increased while
copper sulde formed on the surface was oxidized. This resulted in
a decrease in the oatability of pyrite with the galvanic current. On
the basis of these results, Huang and Grano suggested that it may
be possible to predict and optimize the oatation behavior of pyrite
through the measurement of the galvanic current during grinding.
Effect of water quality
The quality of water used in grinding is another important
factor affecting the galvanic interaction and consequently otation
of pyrite. Moslemi et al. [128] studied effect of presence of sulfate
and chloride ions in water on the rest potentials of pyrite and
grinding media. They found that the presence of these ions causes
an increase in the potential difference between pyrite and grinding
media, resulting in an increase in the galvanic interaction during
grinding.
Effect of reactions occurred in the grinding system
It peroxide hydrogen which is a strong oxidizing agent
[246,264] can greatly reduce oatability of pyrite [265,266], even
in the presence of xanthate collectors [206], by enhancing the
surface oxidation of pyrite and also by oxidizing xanthates to form
perxanthates [199,267,268] which do not have collector properties
[267,269,270]. Studies have shown that hydrogen peroxide can be
produced during grinding of pyrite [271276] both in the presence
[277282] and absence of oxygen [278282]. In the presence of
oxygen, superoxide anion, which is generated by ferrous iron,
reacts further with the ferrous iron to form hydrogen peroxide
(Reactions (34) and (35)) [246,264,278].
Fe2+ + O2 = Fe3+ + (O2)

(34)

Fe2+ + (O2) + 2H+ = Fe3+ + H2O2

(35)

In the absence of oxygen, hydrogen peroxide is formed by


combining two hydroxyl radicals which are generated by reacting
ferric iron with water (Reactions (36) and (37)) [246,264,278].
Fe3+ + H2O = HO + H+ + Fe2+

(36)

2HO = H2O2

(37)

Nooshabadi et al. [246] studied some factors affecting the


production of hydrogen peroxide during grinding and found that
the amount of formed hydrogen peroxide increases with increasing grinding time and solids concentration due to the increased
surface area of solids in contact with water. In addition, they found
that an increase in the electrochemical activity of grinding media
[246,281] as well as a decrease in the pH of the pulp [281,282] leads
to an increase in the formation of hydrogen peroxide because of the

increased production of ferrous iron. The results also showed that


more hydrogen peroxide was produced during grinding under a
nitrogen atmosphere than under an air atmosphere, which was
explained to be due to a change in Eh and pH of the pulp in the
nitrogen atmosphere, resulting in producing more ferrous iron
[246].
Summary and recommendations
In this review, the various electrochemical processes associated
with pyrite occurring in a otation system as well as the key factors
affecting them have been discussed. Pyrite as the most abundant
sulde mineral on the Earth has little economic value and in
otation processes it is often rejected to tailings, unless it contains
signicant amounts of valuable metals to be recovered.
Over several decades, extensive studies have been conducted on
the electrochemical behavior of pyrite in the otation process to
provide the basic knowledge required for optimizing the process.
The results of these studies need to be integrated with each other
to be more effectively applicable to the mineral processing plants.
Due to the electrochemical nature of pyrite, the studies show
that any change in the electrochemical potential may have
signicant effects on the pyrite otation results, whether pyrite
is a gangue or a desired mineral. Table 4 summarizes the
relationship between the electrochemical conditions and processes related to pyrite otation.
These electrochemical processes are affected by various factors
such as pulp pH, pulp solid percent, particle size distribution, pulp
oxygen content, conditioning and otation time, type and
concentration of reducing or oxidizing agent and electrochemical
properties of collector.
Grinding also has signicant effects on these processes and
therefore it is necessary to pay special attention to it. Type of mill,
grinding time, grinding atmosphere, surface area of solid particles,
electrochemical activity of grinding media and also type of
grinding method (dry or wet) are important factors associated
with the grinding process, affecting the electrochemical interactions.
The presence of various impurities in the water used in the
otation, which is probably due to dissolution of various minerals
and also due to otation reagents remaining in recycled water may
change the intensity of the electrochemical interactions and result
in unwanted depression or otation of pyrite. Therefore, the
quality of water used in the otation circuit should be regularly
monitored and maintained at a optimum level.
It is important to consider the geographical origin of pyrites.
Pyrites from different geographical regions, due to having different
chemical and physical properties such as iron to sulfur ratio, crystal
structure and surface morphology exhibit different electrochemical properties, which may result in a various oxidation, depression,
activation, collector adsorption and otation behaviors.
It is well known that the electrochemical interactions do not
occur uniformly throughout the pyrite surface and their intensity
may vary signicantly even between adjacent points. This can lead
to the formation of different products on the surface of different
particles. Therefore, it is important to conduct a comprehensive
study to fully identify the factors causing such variations in the

Please cite this article in press as: H. Moslemi, M. Gharabaghi, A review on electrochemical behavior of pyrite in the froth otation process, J.
Ind. Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.12.012

G Model
JIEC 3218 No. of Pages 18

H. Moslemi, M. Gharabaghi / Journal of Industrial and Engineering Chemistry xxx (2016) xxxxxx

15

Table 4
Relationship between electrochemical conditions and processes in pyrite otation.
Electrochemical process

Electrochemical conditions

Explanation

Flotation of pyrite by
oxidation

Moderately oxidizing potentials are favorable.

 At these potential values, hydrophobic sulfur-rich species are produced on the


surface.
 At these potential values, iron species produced by the oxidation are soluble.

Depression of pyrite by
oxidation

Strongly oxidizing potentials are favorable.

 In these conditions, hydrophobic sulfur species are oxidized into hydrophilic


species.
 In these conditions, hydrophilic iron oxide/hydroxide species are formed and
precipitated on the surface.

Strongly reducing potentials are favorable.

 At strongly reducing potentials, hydrophobic sulfur species present on the surface


are reduced to hydrophilic species.

Activation of pyrite by copper Increasing potential has an adverse effect.

 With Increasing potential, the oxidation rate Cu(I) to Cu(II) increases, and as a result
the concentration of Cu(I) ions which are responsible for the activation decreases.
 At high potentials, since extensive oxidation of the surface occurs, the surface is
covered by iron oxides/hydroxides. Thus, the available surface area for the copper
adsorption decreases.
 At high potentials, hydrophilic copper carbonate/hydroxyl species are formed and
precipitated on the Cu activated surface.

Xanthate adsorption on non- A great decrease or increase in the potential has  A great decrease in potential leads to a decrease in the xanthate oxidation to
activated surface of pyrite adverse effects. The potential must be optimized.
dixanthogen.
 A great increase in potential leads to an increase in the surface oxidation and the
formation of iron oxide/hydroxide species on the surface. These species, in addition
to being hydrophilic, prevent electron transfer across the mineralsolution
interface, and thus, reduce the xanthate adsorption on the surface.

A great decrease or increase in the potential has  A great decrease in potential leads to a decrease in the xanthate oxidation to
Xanthate adsorption on
copper activated surface of adverse effects. The potential must be optimized.
dixanthogen.
 A great increase in potential leads to the formation of copper carbonate/hydroxyl
pyrite
species, in addition to iron oxide/hydroxide species, which are precipitated on the
Cu activated surface. These species, in addition to being hydrophilic, prevent
electron transfer across the mineralsolution interface, and thus, reduce the
xanthate adsorption on the activated surface.

Depression of pyrite by
depressants

Oxidizing potentials are favorable.

intensity of the interactions and to propose more detailed


mechanisms for these electrochemical interactions. Moreover, it
is necessary to accurately study the evolution and the various steps
of formation of the nal products and also the intermediate
products formed at each step in order to provide the possibility of
more precise control, and therefore further optimization of the
processes.
It is essential to note that in order for the experimental results
to be effectively applicable to mineral processing plants, the
experiments must be carried out under conditions as close as
possible to those existing in the plants. For this purpose, the quality
of water and also the type and the dosage of reagents used in the
experiments should be similar to those used in plants. Moreover, in
practice, because pyrite is associated with other sulde minerals,
the electrochemical processes related to the otation are affected
by the galvanic interactions. It is important to consider the effects
of the presence and relative proportions of these minerals in the
experimental studies.
References
[1] C.L. Wiersma, J.D. Rimstidt, Geochim. Cosmochim. Acta 48 (1984) 85.
[2] G.U. Von Oertzen, W.M. Skinner, H.W. Nesbitt, A.R. Pratt, A.N. Buckley, Surf.
Sci. 601 (2007) 5794.
[3] O. Bicak, Z. Ekmekci, D.J. Bradshaw, P.J. Harris, Miner. Eng. 20 (2007) 996.

 In the case of cyanide, it has been found that it is an effective depressant only when
the surface is oxidized.
 In the case of organic depressants, it has been found that increasing potential and
increasing oxidation products density on the surface is often favorable for
depressants adsorption.

[4] A.P. Chandra, A.R. Gerson, Surf. Sci. Rep. 65 (2010) 293.
[5] A. Ceylan, G. Bulut, F. Gktepe, J. Mater. Sci. Eng. B 4 (2014) 119.
[6] Y. Mikhlin, Y. Tomashevich, S. Vorobyev, S. Saikova, A. Romanchenko, R. Felix,
Appl. Surf. Sci. 387 (2016) 796.
[7] J.B. Hiskey, W.J. Schlitt, Proceedings of the second SME-SPE International
Solution Mining Symposium, USA, 1982, pp. 55.
[8] M. Mitovski Aleksandra, N. Mihajlovic Ivan, D. Strbac Nada, D. Sokic Miroslav,
T. Zivkovic Dragana, D. Zivkovic Zivan, Hem. Ind. 69 (2015) 287.
[9] R. Bruce, H. Kadereit, K. Mayhew, R. Mean, A. Nagy, O. Wagner, Proceedings of
the 6th International Seminar on Copper Hydrometallurgy, Vina del Mar,
Chile, 2011.
[10] R. Padilla, C.A. Rivas, M.C. Ruiz, Metall. Mater. Trans. B 39 (2008) 399.
[11] M.C. Ruiz, M.V. Vera, R. Padilla, Hydrometallurgy 105 (2011) 290.
[12] R. Liu, Z. Yang, Z. He, L. Wu, C. Hu, J. Qu, Chem. Eng. J. 304 (2016) 986.
[13] T. Ctvrtnickova, M.P. Mateo, A. Yanez, G. Nicolas, Appl. Surf. Sci. 257 (2011)
5447.
[14] R. Murphy, D.R. Strongin, Surf. Sci. Rep. 64 (2009) 1.
[15] E.S. Rubin, M.A. Cushey, R.J. Marnicio, C.N. Bloyd, J.F. Skea, Environ. Sci.
Technol. 20 (1986) 960.
[16] A. Demirbas, M. Balat, Energy Sources 26 (2004) 541.
[17] M. Flues, P. Hama, M.J.L. Lemes, E.S.K. Dantas, A. Fornaro, Atmos. Environ. 36
(2002) 2397.
[18] P.R. Salve, T. Gobre, H. Lohkare, R.J. Krupadam, A. Bansiwal, D.S. Ramteke, S.R.
Wate, J. Atmos. Chem. 68 (2011) 183.
[19] R.A. Pandey, V.K. Raman, S.Y. Bodkhe, B.K. Handa, A.S. Bal, Fuel 84 (2005) 81.
[20] P.W. Seo, I. Ahmed, S.H. Jhung, Chem. Eng. J. 299 (2016) 236.
[21] R. Tailor, A. Sayari, Chem. Eng. J. 289 (2016) 142.
[22] N.E. Altun, T. Gler, U. Akdemir, Proceedings of the XIIth International
Mineral Processing Symposium, Cappadocia, Turkey, 2010, pp. 295.
[23] P.K. Abraitis, R.A.D. Pattrick, D.J. Vaughan, Int. J. Miner. Process. 74 (2004)
41.

Please cite this article in press as: H. Moslemi, M. Gharabaghi, A review on electrochemical behavior of pyrite in the froth otation process, J.
Ind. Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.12.012

G Model
JIEC 3218 No. of Pages 18

16

H. Moslemi, M. Gharabaghi / Journal of Industrial and Engineering Chemistry xxx (2016) xxxxxx

[24] Y.H. Sung, J. Brugger, C.L. Ciobanu, A. Pring, W. Skinner, M. Nugus, Miner.
Deposita 44 (2009) 765.
[25] J. Chen, Y. Chen, Z. Wei, F. Liu, Miner. Eng. 23 (2010) 1070.
[26] D. Paktunc, D. Kingston, A. Pratt, J. McMullen, Can. Mineral. 44 (2006) 213.
[27] V.A. Chanturia, T.N. Matveeva, T.A. Ivanova, N.K. Gromova, L.B. Lantsova, J.
Min. Sci. 47 (2011) 102.
[28] A.T. Makanza, M.K.G. Vermaak, J.C. Davidtz, Int. J. Miner. Process. 86 (2008)
85.
[29] Y.A. Attia, M. El-Zeky, Hydrometallurgy 22 (1989) 291.
[30] K.A. Kydros, K.A. Matis, I.N. Papadoyannis, P. Mavros, Int. J. Miner. Process. 38
(1993) 141.
[31] A.N. Buckley, R. Woods, Appl. Surf. Sci. 27 (1987) 437.
[32] R.S.C. Smart, J. Amarantidis, W.M. Skinner, C.A. Prestidge, L. LaVanier, S.R.
Grano, Scanning Microsc. 12 (1998) 553.
[33] J. Choi, E. Lee, S.Q. Choi, S. Lee, Y. Han, H. Kim, Chem. Eng. J. 285 (2016) 207.
[34] L.R.G. Santos, A.F. Barbosa, A.D. Souza, V.A. Leao, Miner. Eng. 19 (2006) 1251.
[35] M. Riekkola-Vanhanen, Miner. Eng. 48 (2013) 2.
[36] R.G. Acres, S.L. Harmer, D.A. Beattie, Miner. Eng. 23 (2010) 928.
[37] G. Genn, R. Morrison, Miner. Eng. 62 (2014) 129.
[38] S. Chander, Int. J. Miner. Process. 33 (1991) 121.
[39] W. Tolley, D. Kotlyar, R. Van Wagoner, Miner. Eng. 9 (1996) 603.
[40] A.M. Buswell, D.J. Bradshaw, P.J. Harris, Z. Ekmekci, Miner. Eng. 15 (2002) 395.
[41] X.M. Yuan, B.I. Palsson, K.S.E. Forssberg, Int. J. Miner. Process. 46 (1996) 155.
[42] W.H. Dickinson, Z. Lewandowski, Biodegradation 9 (1998) 11.
[43] R. Herrera-Urbina, F.J. Sotillo, D.W. Fuerstenau, Int. J. Miner. Process. 55
(1998) 113.
[44] S.C. Bouffard, B.F. Rivera-Vasquez, D.G. Dixon, Hydrometallurgy 84 (2006)
225.
[45] S. He, D. Fornasiero, W. Skinner, Miner. Eng. 18 (2005) 1208.
[46] C.L. Caldeira, V.S.T. Ciminelli, A. Dias, K. Osseo-Asare, Int. J. Miner. Process. 72
(2003) 373.
[47] T. Gler, J. Colloid Interface Sci. 288 (2005) 319.
[48] C. Hicyilmaz, N.E. Altun, Z. Ekmekci, G. Gokagac, Colloids Surf. A 233 (2004)
11.
[49] C. Hicyilmaz, N.E. Altun, Z. Ekmekci, G. Gokagac, Miner. Eng. 17 (2004) 879.
[50] N.D. Janetski, S.I. Woodburn, R. Woods, Int. J. Miner. Process. 4 (1977) 227.
[51] G.H. Kelsall, Q. Yin, D.J. Vaughan, K.E.R. England, N.P. Brandon, J. Electroanal.
Chem. 471 (1999) 116.
[52] D. Kocabag, G.H. Kelsall, H.L. Shergold, Int. J. Miner. Process. 29 (1990) 211.
[53] M.B.M. Monte, A.J.B. Dutra, C.R.F. Albuquerque Jr., L.A. Tondo, F.F. Lins, Miner.
Eng. 15 (2002) 1113.
[54] D.P. Tao, P.E. Richardson, G.H. Luttrell, R.H. Yoon, Electrochim. Acta 48 (2003)
3615.
[55] X.H. Wang, J. Colloid Interface Sci. 178 (1996) 628.
[56] G.L. Simmons, 1997 SME Annual Meeting, Denver, USA, February 2427, 1997.
[57] I.C. Hamilton, R. Woods, J. Electroanal. Chem. 118 (1981) 327.
[58] D. Kocabag, T. Gler, Int. J. Miner. Process. 87 (2008) 51.
[59] S.R. Rao, J.A. Finch, Can. Metall. Q. 27 (1988) 253.
[60] N.O. Egiebor, B. Oni, J. Chem. Eng. 2 (2007) 47.
[61] R.D. Marco, S. Bailey, Z. Jiang, J. Morton, R. Chester, Electrochem. Commun. 8
(2006) 1661.
[62] E.T. Pecina, A. Uribe, J.A. Finch, F. Nava, Miner. Eng. 19 (2006) 904.
[63] Z. Ekmekci, H. Demirel, Int. J. Miner. Process. 52 (1997) 31.
[64] P. Donato, M. Kongolo, O. Barres, J. Yvon, F. Enderle, E. Bouquer, M. Alnot, J.M.
Cases, Powder Technol. 105 (1999) 141.
[65] A.N. Buckley, I.C. Hamilton, R. Woods, Flotation of Sulphide Minerals
Conference, Stockholm, Sweden, 1984, pp. 41.
[66] A.N. Buckley, I.C. Hamilton, R. Woods, International Symposium on
Electrochemistry in Mineral and Metal Processing II, Atlanta, USA, May
1520, 1988, pp. 234.
[67] A.N. Buckley, R. Woods, International Symposium on Electrochemistry in
Mineral and Metal Processing, Cincinnati, USA, May 611, 1984, pp. 286.
[68] A.N. Buckley, R. Woods, H.J. Wouterlood, International Symposium on
Electrochemistry in Mineral and Metal Processing II, Atlanta, USA, May 15
20, 1988, pp. 211.
[69] M. Descostes, P. Vitorge, C. Beaucaire, Geochim. Cosmochim. Acta 68 (2004)
4559.
[70] I.V. Chernyshova, J. Electroanal. Chem. 558 (2003) 83.
[71] M.P. Janzen, R.V. Nicholson, J.M. Schaper, Geochim. Cosmochim. Acta 64
(1999) 1511.
[72] H. Moslemi, P. Shamsi, M. Alimohammady, J. South. Afr. Inst. Min. Metall. 112
(2012) 883.
[73] M.M. McGuire, K.N. Jallad, D. Ben-Amotz, R.J. Hamers, Appl. Surf. Sci. 178
(2001) 105.
[74] D.J. Rimstidt, D.J. Vaughan, Geochim. Cosmochim. Acta 67 (2003) 873.
[75] S. Chander, A. Briceno, Int. J. Miner. Process. 24 (1988) 73.
[76] S. Chander, J. Pang, A. Briceno, International Symposium on Electrochemistry
in Mineral and Metal Processing II, Atlanta, USA, May 1520, 1988, pp. 247.
[77] R.H. Yoon, Int. J. Miner. Process. 8 (1981) 31.
[78] G. Luttrel, R.H. Yoon, Int. J. Miner. Process. 13 (1984) 271.
[79] X. Zhu, M.E. Wadsworth, D.M. Bodily, A.M. Riley, Proceedings of the Fourth
International Conference on Processing and Utilization of High Sulfur Coals
IV, Idaho falls, USA, August 2630, 1991, pp. 205.
[80] X. Zhu, J. Li, D.M. Bodily, M.E. Wadsworth, Proceedings of the International
Symposium on Electrochemistry in Mineral and Metal Processing III, Saint
Louis, USA, May 1721, 1992, pp. 391.

[81] M.M. Antonijevic, M.D. Dimitrijevic, S.M. Serbula, V.L.J. Dimitrijevic, G.D.
Bogdanovic, S.M. Milic, Electrochim. Acta 50 (2005) 4160.
[82] M. Descostes, F. Mercier, C. Beaucaire, P. Zuddas, P. Trocellier, Nucl. Instrum.
Methods Phys. Res. Sect. B 181 (2001) 603.
[83] S. Chander, A. Briceno, Miner. Metall. Process. 4 (1987) 171.
[84] D. Kocabag, H.L, G.H. Kelsall, Int. J. Miner. Process. 29 (1990) 211.
[85] T. Pecina, A. Uribe, F. Nava, J.A. Finch, Miner. Eng. 19 (2006) 172.
[86] N.P. Finkelstein, S.A. Allison, V.M. Lovell, B.V. Stewart, Advances in Interfacial
Phenomena of Particulate/Solution/Gas System: Applications to Flotation
Research, AIChE Symposium Series, Salt Lake City, USA, August 2021, 1974,
pp. 165.
[87] G.W. Heyes, W.J. Trahar, Int. J. Miner. Process. 4 (1977) 317.
[88] J.R. Gardner, R. Woods, Int. J. Miner. Process. 6 (1979) 1.
[89] D.P. Tao, Y.Q. Li, P.E. Richardson, R.H. Yoon, Colloids Surf. A 93 (1994) 229.
[90] R.H. Yoon, D.P. Tao, M.X. Lu, P.E. Richardson, G.H. Luttrell, Coal Prep. 18 (1997)
53.
[91] T. Guler, K. Sahbudak, S. Cetinkaya, U. Akdemir, Trans. Nonferrous Met. Soc.
China 23 (2013) 2766.
[92] R. Woods, Int. J. Miner. Process. 72 (2003) 151.
[93] Y. Peng, S. Grano, D. Fornasiero, J. Ralston, Int. J. Miner. Process. 70 (2003) 67.
[94] A.N. Buckley, K.W. Riley, Surf. Interface Anal. 17 (1991) 655.
[95] J.B. Zachwieja, J.J. McCarron, G.W. Walker, A.N. Buckley, J. Colloid Interface
Sci. 132 (1989) 462.
[96] A.N. Buckley, R. Woods, Aust. J. Chem. 37 (1984) 2403.
[97] S. Karthe, R. Szargan, E. Suoninen, Appl. Surf. Sci. 72 (1993) 157.
[98] V. Toniazzo, C. Mustin, J.M. Portal, B. Humbert, R. Benoit, R. Erre, Appl. Surf.
Sci. 143 (1999) 229.
[99] K. Duran, T. Guler, Miner. Eng. 20 (2007) 1246.
[100] G.W. Walker, C.P. Walters, P.E. Richardson, Int. J. Miner. Process. 18 (1986) 119.
[101] A.M. Raichur, X.H. Wang, B.K. Parekh, Miner. Eng. 14 (2001) 65.
[102] X.H. Wang, C.L. Jiang, A.M. Raichur, B.K. Parekh, J.W. Leonard, Proceedings of
the International Symposium on Electrochemistry in Mineral and Metal
Processing III, Saint Louis, USA, May 1721, 1992, pp. 410.
[103] S.U.M. Khan, J.P. Baltrus, R.W. Lai, A.G. Richardson, Appl. Surf. Sci. 47 (1991)
355.
[104] J.E. Gebhart, N.F. Dewsnap, P.E. Richardson, US. Bur. Mines Rep. Invest. (1985)
1.
[105] V.A. Chandurya, V.E. Vigdergauz, M. Teplyakova, Sov. Surf. Eng. Appl.
Electrochem. 21 (1988) 30.
[106] Y. Xian, S. Wen, X. Chen, J. Deng, J. Liu, Int. J. Miner. Metall. Mater. 19 (2012)
1069.
[107] A.P. Deditius, S. Utsunomiya, D. Renock, R.C. Ewing, C.V. Ramana, U. Becker, S.
E. Kesler, Geochim. Cosmochim. Acta 72 (2008) 2919.
[108] A.P. Deditius, S. Utsunomiya, M. Reich, S.E. Kesler, R.C. Ewing, R.M. Hough, J.L.
Walshe, Ore Geol. Rev. 42 (2011) 32.
[109] K.S. Savage, T.N. Tingle, P.A. ODay, G.A. Waychunas, D.K. Bird, Appl. Geochem.
15 (2000) 1219.
[110] O.L. Gaskova, E.V. Belogub, D.V. Makarov, Russ. Geol. Geophys. 51 (2010) 176.
[111] M. Reich, S.E. Kesler, S. Utsunomiya, C.S. Palenik, S.L. Chryssoulis, R.C. Ewing,
Geochim. Cosmochim. Acta 69 (2005) 2781.
[112] A.P. Deditius, S.E. Kesler, R.C. Ewing, S. Utsunomiya, J. Walshe, Proceedings of
the 10th Biennial SGA Meeting of the Society for Geology Applied to Mineral
Deposits, Townsville, Australia, 2009, pp. 710.
[113] K.S. Savage, D. Stefan, S.W. Lehner, Appl. Geochem. 23 (2008) 103.
[114] S. Lehner, K. Savage, M. Ciobanu, D.E. Cliffel, Geochim. Cosmochim. Acta 71
(2007) 2491.
[115] S.W. Lehner, K.S. Savage, J.C. Ayers, J. Cryst. Growth 286 (2006) 306.
[116] Y.Q. Li, J.H. Chen, Y. Chen, Acta Phys. Chim. Sin. 26 (2010) 1435.
[117] R.M. Garrels, C.L. Christ, Solutions, Minerals and Equilibria, First ed., Harper
and Row, New York, 1965.
[118] S. Chander, Miner. Metall. Process. 5 (1988) 104.
[119] E. Peters, the fourth Australian Electrochemistry Conference, Australia,
February 1620, 1976, pp. 267.
[120] R. Woods, P.E. Richardson, Advances in Mineral Processing: a Half-Century of
Progress in Application of Theory to Practice, Proceedings of a Symposium
honouring Nathaniel Arbuten on his 75th birthday, New Orleans, USA, March
35, 1986, pp. 154.
[121] M. Pourbaix, Atlas of Electrochemical Equilibria in Aqueous Solutions, First
ed., Pergamon press, London, 1966.
[122] Y. Peng, B. Wang, A. Gerson, Int. J. Miner. Process. 102103 (2012) 141.
[123] E. Forssberg, T.V. Subrahmanyam, XVIII International Mineral Processing
Congress, Sydney, Australia, May 2328, 1993, pp. 1.
[124] P.J. Guy, W.J. Trahar, Int. J. Miner. Process. 12 (1984) 15.
[125] Z. Wei, W.J. Xu, S.T. Zhong, Z.M. Zong, J. China Univ. Min. Technol. 18 (2008)
0571.
[126] W. Zhao, H. Zhu, Z.M. Zong, J.H. Xia, X.Y. Wei, Fuel 84 (2005) 235.
[127] W.Z. Shen, D. Fornasiero, J. Ralston, Miner. Eng. 11 (1998) 145.
[128] H. Moslemi, P. Shamsi, F. Habashi, Miner. Eng. 24 (2011) 1038.
[129] W.J. Trahar, G.D. Senior, L.K. Shannon, Int. J. Miner. Process. 40 (1994)
287.
[130] C.L. Jiang, X.H. Wang, B.K. Parekh, J.W. Leonard, Colloids Surf. A 136 (1998) 51.
[131] X. Wang, E. Forssberg, J. Colloid Interface Sci. 140 (1990) 217.
[132] Z. Ekmekqi, H. Demirel, Proceedings of the 6th International Mineral
Processing Symposium, Kusadasi, Turkey, September, 2426, 1996, pp. 197.
[133] Y. Cai, Y. Pan, J. Xue, G. Su, Appl. Surf. Sci. 255 (2009) 4066.
[134] Y. Cai, Y. Pan, J. Xue, Q. Sun, G. Su, X. Li, Appl. Surf. Sci. 255 (2009) 8750.

Please cite this article in press as: H. Moslemi, M. Gharabaghi, A review on electrochemical behavior of pyrite in the froth otation process, J.
Ind. Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.12.012

G Model
JIEC 3218 No. of Pages 18

H. Moslemi, M. Gharabaghi / Journal of Industrial and Engineering Chemistry xxx (2016) xxxxxx
[135] J. Leppinen, K. Laajalehto, I. Kartio, E. Suoninen, Proceedings of the 19th
International Mineral Processing Congress, San Francisco, USA, October 22
27, 1995, pp. 35.
[136] J. Pang, S. Chander, Proceedings of the XVIII International Mineral Processing
Congress, Sydney, Australia, May 2328, 1993, pp. 669.
[137] S.R. Grano, H. Cnossen, W. Skinner, W.A. Pretidge, J. Ralston, Int. J. Miner.
Process. 50 (1997) 27.
[138] U. Atalay, C. Hicyilmaz, B. Oztas, Z. Ekmekci, G. Gokagac, Extended Abstracts
of the IXth International Mineral Processing Symposium, Capadoccia, Turkey,
September 1820, 2002, pp. 89.
[139] C. Owusu, J. Addai-Mensah, D. Fornasiero, M. Zanin, Adv. Powder Technol. 24
(2013) 801.
[140] V.V. Hintikka, J.O. Leppinen, Miner. Eng. 8 (1995) 1151.
[141] M. Dimitrijevic, M.M. Antonijevic, V. Dimitrijevic, Miner. Eng. 12 (1999) 165.
[142] J. Pang, S. Chander, Proceedings of the International Symposium on
Electrochemistry in Mineral and Metal Processing III, Saint Louis, USA,
May 1721, 1992, pp. 221.
[143] E. Ahlberg, K.S.E. Forssberg, X. Wang, J. Appl. Electrochem. 20 (1990) 1033.
[144] D. Fornasiero, V. Eijt, J. Ralston, Colloids Surf. 62 (1992) 63.
[145] N.P. Finkelstein, Int. J. Miner. Process. 52 (1997) 81.
[146] A.P. Chandra, A.R. Gerson, Adv. Colloid Interface Sci. 145 (2009) 97.
[147] Y. Peng, S.R. Grano, D. Fornasiero, J. Ralston, Int. J. Miner. Process. 69 (2003)
87.
[148] C. Weisener, A. Gerson, Miner. Eng. 13 (2000) 1329.
[149] C. Weisener, A. Gerson, Surf. Interface Anal. 30 (2000) 454.
[150] A.P. Chandra, L. Puskar, D.J. Simpson, A.R. Gerson, Int. J. Miner. Process. 114
117 (2012) 16.
[151] A.G. Schaufuss, H.W. Nesbitt, I. Kartio, K. Laajalehto, G.M. Bancroft, R. Szargan,
J. Electron Spectrosc. Relat. Phenom. 96 (1998) 69.
[152] A.G. Schaufuss, H.W. Nesbitt, I. Kartio, K. Laajalehto, G.M. Bancroft, R. Szargan,
Surf. Sci. 411 (1998) 321.
[153] K. Laajalehto, J. Leppinen, I. Kartio, T. Laiho, Colloids Surf. A 154 (1999) 193.
[154] Y. Peng, S. Grano, Miner. Eng. 23 (2010) 600.
[155] C.H.G. Bushell, C.J. Krauss, Can. Min. Metall. Bull. 65 (1962) 185.
[156] R. Szargan, S. Karthe, E. Suoninen, Appl. Surf. Sci. 55 (1992) 227.
[157] A.N. Buckley, R. Woods, H.J. Wouterlood, Int. J. Miner. Process. 26 (1989) 29.
[158] X.H. Wang, E. Forssberg, N.J. Bolin, Scand. J. Metall. 18 (1989) 262.
[159] J.O. Leppinen, Int. J. Miner. Process. 30 (1990) 245.
[160] A. Boulton, D. Fornasiero, J. Ralston, Int. J. Miner. Process. 70 (2003) 205.
[161] P.E. Richardson, Z. Chen, D.P. Tao, R.H. Yoon, Proceedings of the 189th
Electrochemical Society Meeting, Los Angeles, USA, May 510, 1996, pp. 179.
[162] S. Voigt, R. Szargan, E. Suoninen, Surf. Interface Anal. 21 (1994) 526.
[163] S.R. Popov, D.R. Vucinic, Colloids Surf. 47 (1990) 81.
[164] Z. Chen, R.H. Yoon, Int. J. Miner. Process. 58 (2000) 57.
[165] W.Z. Shen, D. Fornasiero, J. Ralston, J. Int. J. Miner. Process. 63 (2001) 17.
[166] D. Fornasiero, F. Li, J. Ralston, R.S.C. Smart, J. Colloid Interface Sci. 164 (1994)
333.
[167] R. Herrera-Urbina, D.W. Fuerstenau, Colloids Surf. A 98 (1995) 25.
[168] T. Guler, C. Hicyilmaz, G. Gokagac, Z. Ekmekci, J. Colloid Interface Sci. 279
(2004) 46.
[169] R. Woods, Principles of Mineral FlotationThe Wark Symposium, Adelaide,
Australia, 1983, pp. 91.
[170] P.E. Richardson, G.W. Walker, 15th International Mineral Processing Congress,
Cannes, France, June 29, 1985, pp. 198.
[171] D. Kocabag, Proceedings of the 5th International Mineral Processing
Symposium, Cappadocia, Tukey, September 68, 1994, pp. 105.
[172] T. Camuzcu, U. Akdemir, T. Guler, 10th International Mineral Processing
Symposium, Izmir, Turkey, October 57, 2004, pp. 333.
[173] J.R. De Wet, G. Hodgkinson, P.C. Pistorius, L.C. Prinsloo, R.F. Sandenbergh,
Miner. Eng. 8 (1995) 1333.
[174] H. Kuopanportti, T. Suorsa, O. Dahl, J. Niinimaki, Int. J. Miner. Process. 59
(2000) 327.
[175] K.C. Pillai, J.O.M. Bockris, J. Electrochem. Soc. 131 (1984) 568.
[176] J.O.M. Bockris, K.C. Pillai, Proceedings of the International Symposium on
Electrochemistry in Mineral and Metal Processing I, Cincinnati, USA, 1984, pp.
112.
[177] S.M. Ahmed, Int. J. Miner. Process. 5 (1978) 163.
[178] J. Leppinen, C.O. Basilio, R.H. Yoon, International Symposium on Electrochemistry in Mineral and Metal Processing II, Atlanta, USA, May 1520, 1988,
pp. 49.
[179] R. Woods, Gaudin (A.M.) Memorial Flotation Symposium, Rapid City, USA,
1976, pp. 298.
[180] M.C. Fuerstenau, Principles of Flotation, in: R.P. King (Ed.), South African
Institute of Mining and Metallurgy, Johannesburg, 1982, pp. 159.
[181] J.R. De Wet, P.C. Pistorius, R.F. Sandenbergh, Int. J. Miner. Process. 49 (1997)
149.
[182] T. Biegler, D.A.J. Rand, R. Woods, J. Electroanal. Chem. Interfacial Electrochim.
60 (1975) 151.
[183] T. Biegler, D.A.J. Rand, R. Woods, The Fourth Australian Electrochemistry
Conference, Australia, February 1620, 1976, pp. 291.
[184] W.J. Trahar, Principles of Mineral FlotationThe Wark Symposium, Adelaide,
Australia, 1983, pp. 117.
[185] X.H. Wang, K.S.E. Forssberg, Int. J. Miner. Process. 33 (1991) 275.
[186] X.H. Wang, J. Colloid Interface Sci. 171 (1995) 413.
[187] H.H. Haung, J.D. Miller, Int. J. Miner. Process. 5 (1978) 241.
[188] X.H. Wang, K.S.E. Forssberg, Miner. Eng. 9 (1996) 527.

[189]
[190]
[191]
[192]
[193]
[194]
[195]
[196]
[197]
[198]

[199]
[200]
[201]
[202]
[203]
[204]
[205]
[206]
[207]
[208]
[209]
[210]
[211]
[212]
[213]
[214]
[215]
[216]
[217]
[218]
[219]
[220]
[221]
[222]
[223]
[224]
[225]
[226]
[227]
[228]
[229]
[230]
[231]
[232]
[233]
[234]
[235]
[236]
[237]
[238]
[239]
[240]
[241]
[242]
[243]
[244]
[245]
[246]
[247]
[248]
[249]
[250]
[251]

17

X.H. Wang, K.S.E. Forssberg, N.J. Bolin, Int. J. Miner. Process. 27 (1989) 1.
Q. Zhang, Z. Xu, V. Bozkurt, J.A. Finch, Int. J. Miner. Process. 52 (1997) 187.
G.D. Senior, W.J. Trahar, Int. J. Miner. Process. 33 (1991) 321.
M.C. Fuerstenau, M.C. Kuhn, D.A. Elgillani, Trans. AIME 241 (1968) 148.
J.R. Gardner, R. Woods, Aust. J. Chem. 30 (1977) 981.
C.A. Prestidge, J. Ralston, R. Smart, Int. J. Miner. Process. 38 (1993) 205.
P.J. Harris, N.P. Finkelstein, Int. J. Miner. Process. 2 (1975) 77.
S.A. Allison, L.A. Goold, M.J. Nicol, A. Granville, Metall. Trans. 3 (1972) 2613.
H. Kuopanportti, On the kinetics of conditioning in the otation of pyrite and
chalcopyrite ores, Ph.D. Thesis, University of Oulu, 1998.
A. Gorken, D.R. Nagaraj, P. Riccio, Proceedings of the International
Symposium on Electrochemistry in Mineral and Metal Processing III, Saint
Louis, USA, May 1721, 1992, pp. 92.
A. Uribe-Salas, T.E. Martinez-Cavazos, F.C. Nava-Alonso, J. Mendez-Nonell, C.
Lara-Valenzuela, Int. J. Miner. Process. 59 (2000) 69.
E.T. Pecina-Trevino, A. Uribe-Salas, F. Nava-Alonso, Miner. Eng. 16 (2003) 359.
A.J. Teague, J.S.J. Van Deventer, C. Swaminathan, Int. J. Miner. Process. 57
(1999) 243.
F. Gktepe, Miner. Process. Extr. Metall. Rev.: Int. J. 32 (2010) 24.
H. Kuopanportti, T. Suorsa, E. Pollanen, Miner. Eng. 10 (1997) 1193.
M.C. Fuerstenau, C.A. Natalie, R.M. Rowe, Int. J. Miner. Process. 29 (1990) 89.
M.C. Fuerstenau, M. Misra, B.R. Palmer, Int. J. Miner. Process. 29 (1990) 111.
M.B.M. Monte, F.F. Lins, J.F. Oliveira, Int. J. Miner. Process. 51 (1997) 255.
M.B.M. Monte, F.F. Lins, J.F. Oliveira, Proceedings of the XXI International
Mineral Processing Congress, Rome, Italy, July 2327, 2000, pp. 131.
L.A. Valdivieso, A.A. Sanchez Lopez, S. Song, Int. J. Miner. Process. 77 (2005)
154.
G.C. Allan, J.T. Woodcock, Miner. Eng. 14 (2001) 931.
C.J. Martin, S.R. Rao, J.A. Finch, M. Leroux, Int. J. Miner. Process. 26 (1989) 95.
J.D. Miller, R. Du Plessis, D.G. Kotylar, X. Zhu, G.L. Simmons, Int. J. Miner.
Process. 67 (2002) 1.
T.N. Matveeva, V.A. Chanturia, N.K. Gromova, L.B. Lantsova, E.V. Koporulina, J.
Min. Sci. 49 (2013) 637.
J.H. Chen, Q.M. Feng, Y.P. Lu, Chin. J. Nonferrous Met. 10 (2000) 426.
J.D. Miller, C.L. Lin, S.S. Chang, Coal Prep. 1 (1984) 21.
C. Zhao, D. Huang, J. Chen, Y. Li, Y. Chen, W. Li, Miner. Eng. 95 (2016) 131.
A.M. Ntemi, An evaluation of the current situation of cyanide waste disposal
and treatment methods, Ph.D Thesis, Free University of Berlin, 2013.
P.R. Meyers, Cyanide degradation by Bacillus pumilus Cl: cellular and
molecular characterization, Ph.D Thesis, University of Cape Town, 1993.
D.E. Barnes, P.J. Wright, S.M. Graham, E.A. Jones-Watson, Geostand. Geoanal.
Res. 24 (2000) 183.
N. Lotter, Cyanide volatilization from gold leaching operations and tailing
storage facilities, M. Sc. Thesis, University of Pretoria, 2005.
B. Guo, Y. Peng, R. Spinosa-Gomez, Miner. Eng. 71 (2015) 194.
D.F. Myers, W.E. Fry, Physiol. Biochem. 68 (1978) 1037.
W.E. Fry, R.L. Miller, Arch. Biochem. Biophys. 161 (1972) 468.
J. Wu, J. Wang, H. Liu, S. He, X. Huang, Water Sci. Technol. 64 (2011) 2274.
S. Golbaz, A. Jonidi Jafari, M. Raee, R. Rezaei Kalantary, Chem. Eng. J. 253
(2014) 251.
M.Z. Mubarok, P.S. Irianto, J. Eng. Technol. Sci. 48 (2016) 276.
D.A. Elgillani, M.C. Fuerstenau, Trans. Soc. Min. Eng. AIME 241 (1968) 437.
B. Ball, R.S. Rickard, Gaudin (A.M.) Memorial Flotation Symposium, Rapid
City, USA, 1976, pp. 458.
M.C. Fuerstenau, J.D. Miller, M.C. Kuhn, Chemistry of Flotation, First ed.,
Society of Mining Engineering, New York, 1985.
J.D. Miller, R. Kappes, G.L. Simmons, K.M. LeVier, Miner. Eng. 19 (2006) 659.
M.E. Hoyack, S. Raghavan, Trans. Inst. Min. Metall. Sect. C 96 (1987) 173.
R.K. Rath, S. Subramanian, T. Pradeep, J. Colloid Interface Sci. 229 (2000) 82.
Y. Mu, Y. Peng, R.A. Lauten, Miner. Eng. 9697 (2016) 143.
A. Lopez Valdivieso, T. Celedon Cervantes, S. Song, A. Robledo Cabrera, J.S.
Laskowski, Miner. Eng. 17 (2004) 1001.
A. Lopez Valdivieso, A.A. Sanchez Lopez, S. Song, H.A. Garcia Martinez, S.
Licon Almada, Can. Metall. Q. 46 (2007) 301.
L. Run-qing, S. Wei, H. Yue-hua, W. Dian-zuo, J. Cent. South Univ. Technol. 16
(2009) 0753.
Y. Mu, Y. Peng, R.A. Lauten, Electrochim. Acta 174 (2015) 133.
Y. Mu, Y. Peng, R.A. Lauten, Miner. Eng. 92 (2016) 37.
Y. Mu, Y. Peng, R.A. Lauten, Miner. Eng. 9697 (2016) 113.
A. Fothergill, J.E. Fothergill, Biochem. J. 116 (1970) 555.
J. Liu, Z. Wang, B. Li, Y. Zhang, Trans. Nonferrous Met. Soc. China 16 (2006)
943.
J.A. Rojas-Chapana, H. Tributsch, Hydrometallurgy 59 (2001) 291.
C.M.A. Brett, A.M.O. Brett, Electrochemistry: Principles, Methods, and
Applications, First ed., Oxford University Press, New York, 1993.
E. Peters, SME-AIME Short Course on Bio Extractive Mining, Colorado, USA,
1970, pp. 46.
C. Jia, D. Wei, W. Liu, C. Han, S. Gao, Y. Wang, Trans. Nonferrous Met. Soc. China
18 (2008) 1247.
Y. Peng, S. Grano, Electrochim. Acta 55 (2010) 5470.
A. Javadi Nooshabadi, A.C. Larsson, K.H. Rao, Miner. Eng. 49 (2013) 128.
K. Adam, I. Iwasaki, Int. J. Miner. Process. 12 (1984) 39.
K.A. Natarajan, I. Iwasaki, Int. J. Miner. Process. 13 (1984) 53.
M.K. Yelloji Rao, K.A. Natarajan, Miner. Metall. Process. 7 (1989) 146.
M.K. Yelloji Rao, K.A. Natarajan, Int. J. Miner. Process. 27 (1989) 279.
I. Iwasaki, K.J. Reid, H.A. Lex, K.A. Smith, Min. Eng. (1983) 1184.

Please cite this article in press as: H. Moslemi, M. Gharabaghi, A review on electrochemical behavior of pyrite in the froth otation process, J.
Ind. Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.12.012

G Model
JIEC 3218 No. of Pages 18

18

H. Moslemi, M. Gharabaghi / Journal of Industrial and Engineering Chemistry xxx (2016) xxxxxx

[252] K.S.E. Forssberg, T.V. Subrahmanyam, L.K. Nilsson, Int. J. Miner. Process. 38
(1993) 157.
[253] H. Cheng, K.A. Smith, I. Iwasaki, Proceeding of the Paul E. Queneau
International Symposium: Extractive Metallurgy of Copper, Nickel and
Cobalt, Denver, USA, February, 1993, pp. 971.
[254] X.M. Yuan, B.I. Palsson, K.S.E. Forssberg, Int. J. Miner. Process. 46 (1996) 181.
[255] C.J. Greet, P. Steinier, Proceedings of the Metallurgical Plant Design and
Operating Strategies Conference, Melbourne, Australia, September 67, 2004,
pp. 319.
[256] C.J. Greet, J. Kinal, P. Steinier, Centenary of Flotation Symposium, AusIMM,
Brisbane, Australia, June 69, 2005, pp. 967.
[257] Y. Wei, R.F. Sandenbergh, Miner. Eng. 20 (2007) 264.
[258] G. Gu, J. Dai, H. Wang, G.Z. Qiu, J. Cent. South Univ. Technol. 11 (2004) 275.
[259] X. Chen, Y. Peng, D. Bradshawa, Int. J. Miner. Process. 125 (2013) 129.
[260] Y. Peng, S. Grano, J. Ralston, D. Fornasiero, Miner. Eng. 15 (2002) 493.
[261] G. Huang, S. Grano, Miner. Eng. 18 (2005) 1152.
[262] M.B. McNeil, D.W. Mohr, Geoarchaeology 8 (1993) 23.
[263] H. Liu, C. Zhang, Calphad 25 (2001) 363.
[264] A. Javadi Nooshabadi, K.H. Rao, Adv. Powder Technol. 25 (2014) 832.
[265] S.H. Castro, L. Baltierra, Proceedings of Electrochemistry in Mineral and Metal
Processing VI, Paris, France, May 1418, 2003, pp. 27.
[266] Y. Shi, D. Fornasiero, Chemeca 2010: Engineering at the Edge, Adelaide,
Australia, September 2629, 2010, pp. 541.

[267] M.H. Jones, J.T. Woodcock, Int. J. Miner. Process. 5 (1978) 285.
[268] X.H. Chen, Y.H. Hu, H. Peng, X.F. Cao, J. Cent. South Univ. 22 (2015) 495.
[269] F. Nava-Alonso, T. Pecina-Trevio, R. Perez-Garibay, A. Uribe-Salas, Can.
Metall. Q. 41 (2002) 391.
[270] D. Fornasiero, M. Montalti, J. Ralston, Colloid Interface Sci. 172 (1995) 467.
[271] M.J. Borda, A.R. Elsetinow, M.A. Schoonen, D.R. Strongin, Astrobiology 1
(2001) 283.
[272] M.J. Borda, A.R. Elsetinow, D.R. Strongin, M.A. Schoonen, Geochim.
Cosmochim. Acta 67 (2003) 935.
[273] C. Cohn, S. Mueller, E. Wimmer, N. Leifer, S. Greenbaum, D. Strongin, M.
Schoonen, Geochem. Trans. 7 (2006) 1.
[274] G. Jones, K. Corin, R. van Hille, S. Harrison, Miner. Eng. 24 (2011) 1198.
[275] G. Jones, R. van Hille, S. Harrison, Appl. Microbiol. Biotechnol. 97 (2013) 2735.
[276] G. Jones, B. Megan, R. van Hille, S. Harrison, Appl. Geochem. 29 (2013) 199.
[277] C.A. Cohn, R. Laffers, M.A.A. Schoonen, Environ. Sci. Technol. 40 (2006) 2838.
[278] A. Javadi Nooshabadi, K.H. Rao, Hydrometallurgy 141 (2014) 82.
[279] A. Javadi Nooshabadi, K.H. Rao, Miner. Metall. Process. 30 (2013) 212.
[280] A. Javadi Nooshabadi, K.H. Rao, Int. J. Miner. Process. 125 (2013) 78.
[281] A. Javadi Nooshabadi, K.H. Rao, Proceedings of the XIII International Seminar
on Mineral Processing Technology, Bhubaneswar, India, December 1012,
2013, pp. 169.
[282] A. Javadi Nooshabadi, K.H. Rao, XXVII International Mineral Processing
Congress (IMPC 2014), Santiago, Chile, October 2024, 2014.

Please cite this article in press as: H. Moslemi, M. Gharabaghi, A review on electrochemical behavior of pyrite in the froth otation process, J.
Ind. Eng. Chem. (2016), http://dx.doi.org/10.1016/j.jiec.2016.12.012

You might also like