Sei sulla pagina 1di 36

 

 
Electrochemical study of chalcopyrite dissolution in sulphuric, nitric and
hydrochloric acid solutions
Tatiana das Chagas Almeida, Eric Marsalha Garcia, Hugo Walisson Alves da
Silva, Tulio Matencio, Vanessa de Freitas Cunha Lins
PII:
DOI:
Reference:

S0301-7516(16)30025-4
doi: 10.1016/j.minpro.2016.02.001
MINPRO 2861

To appear in:

International Journal of Mineral Processing

Received date:
Revised date:
Accepted date:

15 November 2014
29 January 2016
2 February 2016

Please cite this article as: Almeida, Tatiana das Chagas, Garcia, Eric Marsalha, da Silva,
Hugo Walisson Alves, Matencio, Tulio, de Freitas Cunha Lins, Vanessa, Electrochemical
study of chalcopyrite dissolution in sulphuric, nitric and hydrochloric acid solutions,
International Journal of Mineral Processing (2016), doi: 10.1016/j.minpro.2016.02.001

This is a PDF le of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its nal form. Please note that during the production process
errors may be discovered which could aect the content, and all legal disclaimers that
apply to the journal pertain.

ACCEPTED MANUSCRIPT
Electrochemical study of chalcopyrite dissolution in sulphuric, nitric
and hydrochloric acid solutions

Tatiana das Chagas Almeidaa, Eric Marsalha Garciab, Hugo Walisson Alves da Silvac,
Tulio Matenciod Vanessa de Freitas Cunha Linse
a

IP

Federal University of Rio de Janeiro, Horcio Macedo Avenue 2030, Zip Code 21941-914, Rio de
Janeiro, Brazil, tatiana das chagas @gmail.com
b

SC
R

Federal University of So Joo del Rey, Frei Orlando 170, So Joo del-Rei, Brazil, Zip Code 36307352, Belo Horizonte, Brazil, ericmgmg@hotmail.com.
c

Federal University of Minas Gerais, Antonio Carlos Avenue 6627, Zip Code 31270901, Belo Horizonte,
Brazil, hugowalisson@gmail.com.
d

NU

Laboratory of Materials and Fuel Cells, Chemistry Department, Federal University of Minas Gerais,
Antonio Carlos Avenue 6627, Belo Horizonte, Brazil, tmatencio@ufmg.br.
e

MA

Corrosion and Surface Engineering Laboratory, Chemical Engineering Department, Federal University
of Minas Gerais, Antonio Carlos Avenue 6627, fax number + 55 31 34091789, phone number + 55 31
34091781, Zip Code 31270901, Belo Horizonte, Brazil, vlins@deq.ufmg.br

TE

Abstract

Electrochemical and surface analyses were carried out to study the leaching of chalcopyrite in acid media,

CE
P

aiming to increase copper extraction from low-grade chalcopyrite ores. Unpublished results include the
use of electrochemical impedance spectroscopy (EIS) to characterize the dissolution resistance of
chalcopyrite surfaces in 0.1 molL-1 of hydrochloric, nitric or sulphuric acids. Potentiodynamic

AC

polarization, atomic absorption spectrometry and EIS analysis showed that hydrochloric acid solutions are
more efficient leaching agents than nitric and sulphuric acids. The impedance results suggested that the
chalcopyrite dissolution is a diffusion-controlled process in hydrochloric and sulphuric acids. The use of
Raman spectroscopy and scanning electron microscopy with energy dispersive spectrometers (SEM/EDS)
allowed the partial identification of lead and bismuth sulphides as impurities. Two products were
identified on the surface of chalcopyrite after anodic polarization, i.e., sulphur in the sulphuric acid only
and covellite in all three acids.
Key words: chalcopyrite leaching; electrochemical impedance spectroscopy; acid dissolution; Raman
spectroscopy

ACCEPTED MANUSCRIPT
1- Introduction

Chalcopyrite (CuFeS2) is the Earths most abundant Cu bearing mineral, and accounts for approximately

70% of the worlds copper reserves (Li et al., 2015, Cordoba et al., 2008). The extraction of copper from

IP

ore may be conducted using two different processes: the pyrometallurgical and the hydrometallurgical.

SC
R

The pyrometallurgical process has dominated the copper industry since the late nineteenth century
(Dreisinger, 2006, Habashi, 2009). However, this scenario is changing slowly. This process is not suitable
for the treatment of complex and low-grade ores, and has high capital for smelting and refining plants, as

NU

well as negative environmental impact caused by the generation of sulphur dioxide (SO2) (Prasad and
Pandey, 1998). In contrast, 18% of copper extraction comes from the hydrometallurgical processes. The

MA

advantages of hydrometallurgy are the ability to treat complex ores with low copper content, small initial
capital expenditure required for the installation of industrial plants, and the viability for small or large
plants (Dutrizac, 1992).

The application of the hydrometallurgical process for the production of copper is promising but has not

TE

yet been consolidated, as there are several challenges to be overcome. The difficulty of employing such a
method for the hydrometallurgical extraction of copper from chalcopyrite is the leaching step, where the

CE
P

copper dissolution from chalcopyrite in an acid medium is slow at low temperatures and tends to decrease
with time (Majuste et al., 2012). This slow kinetics is attributed to the formation of a protective layer on
the mineral surface, inhibiting continued dissolution (Crdoba et al., 2008). The hindered chalcopyrite

AC

dissolution process has also been reported elsewhere (Majuste et al., 2013).
Despite many years of studies, the formation of this layer and its characteristics are not yet fully
established. Few studies have used the technique of electrochemical impedance spectroscopy (EIS),
which is an important tool for the study of surfaces (Hiroyoshi et al., 2004, Dakubo et al., 2012,
Ghahremaninezhad et al., 2012). Hiroyoshi et al. (2004) studied the synergistic effect of cupric and
ferrous ions on active-passive behavior during anodic dissolution of chalcopyrite in sulphuric acid
solutions, using EIS technique. They reported that a high-resistance passive layer grows on the
chalcopyrite surface without cupric and/or ferrous ions, and that coexistence of these ions causes the
formation of another product layer and inhibits the formation of intermediate Cu2S in the active region.
Dakubo et al. (2012) investigated the effectiveness of peroxodisulphate for enhancing copper leaching
rates from chalcopyrite, using sulphuric acid leaching solutions at pH 2. The presence of

ACCEPTED MANUSCRIPT
peroxodisulphate greatly increased copper leaching rates. Gharehmaninezhad et al. (2012) evaluated the
surface of chalcopyrite during dissolution in 0.5 molL-1 sulphuric acid solutions using electrochemical
techniques. EIS analysis at open circuit potential proved the existence of a thin surface layer on the

electrode. This layer was stable up to 100 mV vs. mercury/mercury sulphate electrode (MSE) and was

IP

assumed to be Cu1-xFe1-yS2. By increasing the potential to the range of 100-300 mV (MSE), the previously

SC
R

formed layer partially dissolved and a second layer (Cu 1-x-zS2) was formed on the surface. Both layers
showed the characteristics of passive layers at a low scan rate (0.05 mVs-1) but acted like pseudo-passive
layers at high scan rates ( 50 mVs-1). However, in the potential range of 300-420 mV (MSE), both

NU

surface layers dissolved and active dissolution of the electrode started (Gharehmaninezhad et al., 2012).
Direct leaching of chalcopyrite concentrate includes the use of sulphuric acid, nitric acid, chloride ions,

MA

oxygen leaching, ammonia systems, alcoholic acid media, and biological processes (Venkatachalam,
1991, Solis-Marcial et al., 2014, Gharehmaninezhad et al., 2010, Li et al., 2013, Wang et al., 2014a,

Wang et al., 2014b, Feng et al., 2014, Watling, 2014, Zhang et al., 2014, Smith and Sohn, 2014,

TE

Abdollahi et al., 2014, Khoshkhoo et al., 2014). The aim of this work was to understand the chalchopyrite
dissolution in hydrochloric, sulphuric and nitric acids through an electrochemical study using anodic

CE
P

potentiodynamic tests, EIS, and Raman spectroscopy for the characterization of the chalcopyrite surface
after dissolution. The electrochemical parameters such as critical current density, passivation current
density, and transpassivation potential were obtained using anodic potentiodynamic polarization tests.

AC

The charge transfer resistance and mechanistic information were obtained using electrochemical
impedance spectroscopy.

2- Material and Methods

The chalcopyrite samples were purchased from Wards Natural Science, Rochester, N.Y. Chalcopyrite
electrodes were prepared as per Majuste et al. (2012). Firstly, massive samples were cut using a diamond
wafering blade (Buehler, n.11-4246), to obtain specimens with 1.0 cm2 of exposed areas and 0.5 cm
thicknesses. The specimens were rinsed with double-distilled water and dried with analytical grade
acetone 100% (Synth). A Cu wire within a glass tube was then attached to each specimen using a
conductive silver paint (Dotite, D-550). The samples were, then, mounted in epoxy resins (Epofix,
Struers); these resin blocks, once dried and hardened, were mechanically wet-polished using SiC papers

ACCEPTED MANUSCRIPT
(grit sizes 1000, 1200 and 2400) and subsequently alumina pastes (1.0 and 0.05 m) to obtain fresh and
smooth electrode surfaces (Majuste et al., 2012). Next, the electrodes were rinsed with double-distilled
water in an ultrasonic bath for 15 min, dried with analytical grade ethyl alcohol 95% (Synth) and kept

under vacuum at room temperature (Majuste et al., 2012). The chalcopyrite samples used were

IP

characterized in a previous work (Majuste et al., 2012).

SC
R

0.1 molL-1 of hydrochloric, nitric and sulphuric acid solutions were prepared with deionized water and
concentrated hydrochloric, nitric and sulphuric acids, all analytical grade reagents (Synth ). Quantitative
analysis of iron and copper contents in the leachates was performed using the atomic absorption

NU

spectrometer Hitachi-Z8200 coupled to a Hitachi graphite furnace. Atomic absorption spectroscopy

MA

analyses of acid solutions were performed before and after polarization tests.

Electrochemical tests were performed in a 150 mL electrochemical cell, using a Versa Stat 2 Princeton
Applied Research potentiostat, and Versa Studio software. All experiments were carried out without

TE

stirring. The reference electrode used was Ag/ AgCl, KCl (sat) and the counter electrode was platinum.
The anodic potentiodynamic polarization and EIS were performed at room temperature 25 - 27 0.05C

CE
P

in 0.1 molL-1 of hydrochloric, nitric and sulphuric acid solutions. The anodic polarization was performed
in the potential range of 0.1 to 1.3 V (Ag/AgCl), at a scan rate of 0.05 Vs-1.

AC

For the EIS analysis, the potential amplitude and frequency range were 10 mV and 105 - 10-2 Hz,
respectively. The treatment and equivalent circuit simulation were carried out using the ZviewTM
software.

The surface of the chalcopyrite electrode was analyzed using a scanning electron microscope (JEOL, JSM
6360 LV) with energy dispersive spectrometers (Thermo Noram, Quest). The analysis of the electrode
surface was also made by Raman spectroscopy using a Jobin Yvon Horiba instrument (Labram HR800)
equipped with a CCD detector and He-Ne laser operating at 633 nm. The power used was 20 mW focused
on the sample by a BX-41 Olympus microscope comprising lenses with magnifications of 10, 50 and
100x. The spectral resolution was 1 cm-1. An average of 10 accumulations in the acquisition time of 60
seconds was performed.

ACCEPTED MANUSCRIPT
The scanning electron microscopy (SEM), the energy dispersive spectrometry (EDS) and Raman
spectroscopy were performed immediately after sample preparation and after anodic polarization up to

1.3 V (Ag/AgCl).

SC
R

IP

3- Results and Discussion

3.1 Anodic potentiodynamic polarization

The anodic polarization curves of the chalcopyrite in three types of acids are shown in Figure 1. The

NU

polarization curves obtained with a high potential range in the three acid media are useful to give a
general idea of the electrochemical behavior of the electrode in the medium studied. A tendency to

MA

concentration polarization (McCafferty, 2010) was observed in the cathodic branch of the curve of
chalcopyrite in the nitric acid. On the cathodic side of the curve in HNO3 medium, increased hydrogen
evolution at large cathodic overvoltage may cause a depletion of protons near the electrode surface. In

addition, the increased number of hydrogen gas molecules produced may be slow to diffuse away from

TE

the electrode surface. Either event will result in concentration polarization (McCafferty, 2010) for the
cathodic reaction which was observed only in the nitric acid. In this medium, at high cathodic

CE
P

overvoltage, an inhibition of the cathodic reaction was identified. When electrolytes are in the sulphuric
and hydrochloric acids, the reduction of protons with hydrogen evolution occurred without inhibition
until 0.1 V (Ag/AgCl).

AC

The electrochemical parameters, obtained using the potentiodynamic polarization technique, are shown in
Table 1. The highest critical current density (McCafferty, 2010) to produce a protective layer was
observed in the case of sulphuric acid. Another advantage of the use of sulphuric acid as a leaching agent
is that the current density at the region of the protective layer is one order of magnitude higher than the
current density at the region of the protective layer produced in the nitric acid. The protective layer
dissolves at a higher rate in sulphuric acid than in nitric acid. However the protective layer breaks at a
higher potential in the sulphuric acid, hindering the leaching process. In the hydrochloric acid, the
dissolution of chalcopyrite is continuous after a very small region of relatively constant current. The
hydrochloric acid is not an oxidizing medium, so does not favor the formation of a protective layer of
corrosion product. The chalcopyrite dissolution is facilitated in hydrochloric acid due to the lowest
transpassivation potential in this medium.

ACCEPTED MANUSCRIPT
The previous reduction of the chalcopyrite surface (Figure 1) contributes to the formation of a passive or
pseudo-passive layer in oxidizing acids such as sulphuric and nitric. By reducing metal ions on the
chalcopyrite surface, metallic iron and metallic copper regions appeared generating a galvanic pair, where

iron is preferably oxidized. The metallic copper area is a cathodic region where the cathodic reactions

IP

occurred:

SC
R

2H+ + 2e- H2
4H+ + O2 + 4e- 2H2O

(1)
(2)

On the increase of the cathode area, a higher consumption of electrons occurs favoring the dissolution of

NU

chalcopyrite.

Atomic absorption spectroscopy results (Table 2) showed that after the anodic polarization up to 1.3 V vs.

MA

Ag/AgCl, higher Cu contents and Cu/Fe ratio were obtained in the sulphuric and hydrochloric acid
leaching. Leaching with the nitric acid resulted in the lowest leaching efficiency relative to the copper
extraction. According to the literature, in sulphuric acids, increase in the potential up to 1.085 V vs. SHE

causes the complete release of iron from the chalcopyrite lattice into solution and the formation of a CuS

TE

layer on the electrode surface according to the reaction (Ghahremaninezhad et al., 2010):
2CuFeS2 + 13H2O 0.75CuS + 1.25Cu2+ + Fe2(SO4)3 + 0.25SO42- + 26H+ + 28e-

(3)

CE
P

The CuS layer is unstable at potentials higher than 1.165 V vs. SHE. Therefore, the complete dissolution
of chalcopyrite occurs by further increasing the potential, according to reactions (Viramontes-Gamboa et
al., 2007):

(4)

2CuFeS2 + 16H2O 2Cu2+ + Fe2(SO4)3 + SO42- + 32H+ + 34e-

(5)

AC

CuS + 4H20Cu2+ + SO42- + 8H+ + 8e-

The concentrations of iron and copper measured using atomic absorption spectroscopy were obtained
after polarization up to 1.3 V vs. Ag/AgCl, which is near the potential range described in the literature for
the extraction of iron and copper from chalcopyrite.
The main anodic reactions proposed in the literature for the dissolution of chalcopyrite in 1M
hydrochloric acid solution are (Gharehmaninezhad et al., 2010):
CuFeS2 Cu2+ + Fe3+ + 2S + 5e-

(6)

CuFeS2 + 8H2O Cu2+ + Fe3+ + 2SO42- + 16H+ + 17e-

(7)

3.2 Electrochemical impedance spectroscopy

ACCEPTED MANUSCRIPT
Measurements of open circuit potential were performed until potential stabilization (Figure 2). The
profiles in Figure 2 were almost the same for the three acids. The curve is ascending, starting at a
potential below 200 mV (Ag/AgCl), except for the first analysis (Test 1) in hydrochloric acid starting at

210 mV (Ag/AgCl). Each test was conducted for a minimum period of 1 h. After 30 min., the potential

IP

stabilized and slightly varied at a rate lesser than 0.3 mV/min. The rapid increase in potential during the

SC
R

first 15 min. of immersion in the acid solutions suggests the formation of a corrosion product on the
electrode surface and an occurrence of ohmic polarization (McCafferty, 2010). Table 3 shows the mixed
potential (open circuit potential) of chalcopyrite obtained for the three acid. The average potential,

NU

standard deviation and variation coefficient were calculated. The standard deviations (SD) of corrosion
potentials are shown in Table 3, with SD being greatest for leaching with hydrochloric acid.

MA

EIS was performed at a potential amplitude of 10 mV (Ag/AgCl) in relation to the stabilized open circuit
potential. The Kramers-Kroning (K-K) relationships were applied to validate experimental results. Any
system that satisfies the priori conditions of linearity, stability, and causality must satisfy the K-K

relationships. The K-K technique transforms the real into the imaginary component and vice versa, so that

TE

the transformed quantities may be compared directly with their corresponding experimental values for the
same parameters. A good agreement between the experimental and transformed impedance data for both

CE
P

real and imaginary components validates the EIS data (Zheng et al., 2014).
The Nyquist diagram of chalcopyrite in the three acids is shown in Figure 3a. The highest impedance was
obtained for the dissolution of chalcopyrite in the sulphuric acid. Dreisinger and Abed (2002) reported

AC

that hydrochloric acid is more efficient than sulphuric acid as the activity of Fe3+ in sulphuric acid
solution is reduced due to the formation of FeSO4+. Nicol et al. (2010) pointed out that the existence of
chloride ions can enhance the chalcopyrite leaching rate. The equivalent circuit fitted to the EIS data is
shown in Figure 4 a for the hydrochloric and sulphuric acids and in Figure 4 b for the nitric acid. The
electrochemical parameters obtained from the EIS measurements are shown in Table 3. As compared to
this work, Yu et al. (2011) reported the same equivalent circuit for the dissolution of chalcopyrite in
electrolyte solution containing ethyl xanthate. The first value at high frequencies (105 Hz) is associated
with the solution resistance (R1) which was very low and similar for the three acids used (Table 3).
At the intermediate frequency (from 100 to 1 Hz), a parallel circuit was identified and a charge transfer
resistance (R2) of a surface film with a constant phase element (CPE1), which is a capacitance of the
surface film, was also observed. R3 is the charge transfer resistance occurring within the pores of the

ACCEPTED MANUSCRIPT
surface film (Yu et al., 2011), and CPE2 is the associated capacitance. CPE is encountered frequently in
solid state electrochemistry, and the CPE behavior of interfaces has been ascribed to a fractal nature
(special geometry of the roughness) of the interface. The admittance representation of CPE is given by:

Y() = Y0(j)n = Y0 n cos(n/2) + jY0 n sin(n/2)

(8)

IP

where is the angular frequency, j is the imaginary unit (-1), Y0 is the admittance of an ideal

SC
R

capacitance, and n is the parameter which characterizes deviation of the system from ideal capacitive
behavior. When n=1, the CPE describes an ideal capacitor corresponding to an ideal flat electrode. When
n=0.5, the CPE represents a Warburg impedance related to the ionic flux across the film, corresponding to

NU

porous electrodes (Rammelt and Reinhard, 1990). In this work, for the CPE in series with R3 (Figure 4a),
n=0.40 (Ws-P in Table 3) for the dissolution of chalcopyrite in hydrochloric acid and n = 0.45 for the

MA

dissolution of chalcoprite in sulphuric acid, the latter of which is considered the Warburg impedance (WsR in Table 3) related to the ionic flux across the surface film. The simulation associated with the
impedance spectra indicates that the process is controlled by diffusion in the product layer in the

hydrochloric and sulphuric acids. The impedance associated with a diffusive control of the dissolution is

TE

W1. The element W1 or Ws is a finite length diffusion element used in the case where one boundary

CE
P

imposes a fixed concentration for the diffusing species, and thus it is permeable to the diffusing species.
This type of dispersion relation is generally found in corrosion related diffusion (Boukamp, 1985). The
dispersion relation contained in the admittance represents a cotangent-hyperbolic function:
(9)

AC

Y() = Y0(j)1/2 coth [B (j)1/2]

The peak in the curve of Bode diagram of phase angle (Figure 3b) is broad and can signify more than one
time constant with similar values. The maximum of the phase angle curve of the chalcopyrite in
hydrochloric acid occurred at a lower angle than the maximum of chalcopyrite in nitric and sulphuric acid
solutions, indicating a more resistive behavior, which is beneficial to the chalcopyrite dissolution. The
curve of phase angle starts around = 38 at low frequencies in the case of hydrochloric acid, and around
= 37 for sulphuric acid. The angle associated with the Warburg element is 45o, when the semi-infinite
linear diffusion occurs. However, other diffusive elements are useful models for diffusion when a thin
film (finite diffusion) is involved. The diffusive elements are closer to the Warburg element for the
hydrochloric and sulphuric acids. Yu et al. (2011) also proposed the existence of Warburg impedance for
the chalcopyrite dissolution. In sulphuric and hydrochloric acids, a diffusion control of the chalcopyrite

ACCEPTED MANUSCRIPT
dissolution process is proposed. As the chalcopyrite in the sulphuric acid showed the highest Warburg
impedance value (Table 3), the chalcopyrite dissolution is disadvantaged in this medium.
Hiroyoshi et al. (2004) used EIS to study the effect of Cu2+ and Fe2+ in the formation of the passive layer

IP

on the surface of chalcopyrite in sulphuric acids. A series circuit between Re (solution resistance) and a
parallel R-CPE circuit (resistance - constant phase element) was assumed as equivalent circuit. Hiroyoshi

SC
R

et al. (2004) emphasized that the final frequency (10-1 Hz, an intermediate frequency) probably prevented
some elements from being observed, such as a diffusion element, since they occur at lower frequencies.

presence of the Fe

2+

/ Fe

3+

NU

Ghahremaninezhad et al. (2012) used EIS to study the kinetics of chalcopyrite leaching process in the
redox couple in sulphuric acids. The authors observed a diffusion control of

MA

the chalcopyrite dissolution and also found that the slow dissolution step of chalcopyrite is due to the
slow diffusion of iron through a passive film.

The chemical leaching of chalcopyrite in nitric acid is complex. It has been postulated that the species

which reacts with sulphide is NO+ rather than NO3-, oxidizing sulphide to sulphur. A by-product is the

TE

NO(g), which further reacts with NO2(aq) and regenerated NO+ in acid media (Anderson et al., 2011). The

CE
P

results of anodic polarization suggested a diffusion control of the cathodic reaction in nitric acids.
3.3 Scanning electron microscopy and dispersive energy spectroscopy

AC

SEM images of chalcopyrite before and after polarization up to 1.3 V (Ag/AgCl) (Figure 1) are shown in
Figures 5, 6 and 7 for leaching in the hydrochloric, nitric and sulphuric acid solutions, respectively. In
Figure 5, the appearance of dark regions, which are deeper than the light regions is an evidence of
localized corrosion for leaching in hydrochloric acid. The EDS results showed a semi-quantitative
characterization of typical chalcopyrite, which consisted primarily of S (34.17 wt.%), Fe (30.67 wt.%)
and Cu (34.97 wt.%). Figure 6 shows an intergranular corrosion of chalcopyrite in the nitric acid solution.
The EDS analysis of the chalcopyrite surface resulted in S (35.10 wt.%), Fe (30.69 wt.%) and Cu (33.95
wt.%). The EDS analysis of chalcopyrite surface before the polarization test in sulphuric acid revealed the
following composition: S (35.02 wt.%), Fe (30.86 wt.%) and Cu (33.91 wt.%).
After polarizations with hydrochloric, nitric and sulphuric acids, the wt.% of Cu, Fe, and S did not change
significantly (with up to 7.5% variation). However, dark regions appeared showing a localized attack of

ACCEPTED MANUSCRIPT
the electrode surface by the leach medium. The SEM analysis identified a localized corrosion of the
electrode surface in all electrolytes, being intergranular in the nitric acid.

IP

3.4 Raman spectroscopy

SC
R

Raman spectroscopy was used to characterize the matrix (chalcopyrite) and impurities on the surface of
chalcopyrite, and especially to elucidate possible reaction products formed by the dissolution of the
electrode after polarization in 0.1 molL-1 hydrochloric, nitric and sulphuric acids. The measurement

NU

depth of Raman microspectroscopic analysis is usually at the m scale, depending on the materials for
analysis (Qian et al., 2015).

MA

Figure 8 is an optical micrograph of the polished surface of chalcopyrite electrode and shows the
spectrum collected from area 1 (a). In the range from 75 cm-1 to 550 cm-1, there is only one intense band
at 294 cm-1 that corresponds to chalcopyrite (Mernagh and Trudu, 1993). The spectrum also exhibits

additional bands at 78, 90, 105, 270, 320, 359 and 377 cm-1, all with relatively low intensities. Bands at

TE

270, 320 and 359 cm-1 are reported in literature as chalcopyrite Raman bands with normalized intensities

CE
P

of 0.10, 0.13 and 0.11 (White, 2009). It is important to note that small differences in the intensity and
position of these bands in the spectra, as discussed in the literature (Majuste et al., 2012, Mernagh and
Trudu, 1993, Parker et al., 2008, White, 2009) can be associated with different experimental conditions,

AC

differences in geological formation conditions of samples, variations in iron, copper and sulphur contents
in the chalcopyrite, and features in its crystalline structure. The bands at 291, 320 and 352 cm-1 were
suggested to be associated with the molecular vibrations of the Fe-S bond, and the 265 cm-1 band was
thought to be from the Cu-S bond (White, 2009).
Figure 9 presents a Raman spectrum of area 1 (b) as shown in the optical graph. This grain with color
contrast in the optical image was the same as that for the EDS analysis of chalcopyrite surface after nitric
acid polarization, and comprised a lead bismuth sulphide phase (16.39 wt.% S), (39.37 wt.% Pb) and
(28.72 wt.% Bi). Similarly, Lu and Dreisinger (2013) reported high contents of impurities including Bi
and Pb in a chalcopyrite concentrate. The Raman spectrum obtained for area 1 (b) is not well defined and
only shows several minor broad bands between 80 and 320 cm-1. This spectrum appears to be similar to
the spectra of lead and bismuth sulphides as mineral galenobismutita (PbBi2S4) and aschamalmita
(Pb6Bi2S9) (RRUFF, 2013).

ACCEPTED MANUSCRIPT
After polarization up to 1.3 V (Ag/AgCl) in the hydrochloric acid, regions with different colors, under the
optical microscope (with Raman spectroscopy) were formed on the chalcopyrite surface (Figure 10). A
strong band of the chalcopyrite at 294 cm-1 is still evident after anodic polarization, and appears with

different intensities in the marked areas. In the yellowish area (area 2), the intensity of this band is higher

IP

than that for other areas. Areas 3 and 4 showed decreases in signal intensity from the 294 cm-1 band

SC
R

typical of chalcopyrite, as compared to that for area 2 and increases in the 471 cm-1 band, suggesting the
growth of a surface film. This 471 cm-1 band did not appear on the polished electrode surface (without the
presence of corrosion products). Although the position of this band is similar to the sulphur band at 472

NU

cm-1, the absence of the band at 219 cm-1, characteristic of sulphur, suggests that the band at 471 cm-1 is
due to the presence of another phase. The Raman spectrum of covellite (CuS) shows bands at wave

MA

numbers 471 and 263 cm-1 (Mernagh and Trudu, 1993) with strong and weak intensities, respectively.
Based on previously published work, it is suggested that covellite may be a corrosion product formed on
the chalcopyrite surface leached in hydrochloric acid.

Raman spectra of the chalcopyrite electrode performed after polarization in the nitric acid were obtained

TE

for the areas with different colors (green, red and purple) as shown in Figure 11. It was found that the
three areas all showed sharp decreases in intensity of the chalcopyrite band at 294 cm-1 and increases in

CE
P

the intensities of the bands in the range of 471 to 475 cm-1. This suggests that a corrosion product,
possibly covellite as discussed above, may have formed, which explains the reduction in the intensity of
the typical chalcopyrite vibrational band at 294 cm-1. Sulphur was not identified after anodic polarization

AC

in nitric acid. Earlier publications have reported that the sulphur formed during the leaching in nitric acid
media is oxidized to sulphate via reactions with NO(g) and oxygen (Prasad and Pandey, 1998,
Venkatachalam, 1991, Anderson et al., 2011).
The optical micrograph of the chalcopyrite electrode after anodic polarization in sulphuric acid showed a
different staining (Figure 12), and the Raman spectra were collected to characterize these distinct areas.
The spectra (Figure 12) for the areas 3 and 4 were the same, suggesting similar surface compositions
within the two areas, despite of the color contrast under the optical microscope. Area 2, however, showed
a quite different spectrum, with additional bands at 472, 219 and 153 cm-1 , as compared to those for areas
(3) and (4). These three new bands from area 2 are typical of sulphur. Elemental sulphur (S8) with
orthorhombic crystal structure shows very strong bands at 219 and 472 cm-1 (Parker et al., 2008, White,
2009). A less intense band can be seen at 153 cm-1, and additional very weak bands at 437, 246 and 187

ACCEPTED MANUSCRIPT
cm-1. The bands in the 100300 cm1 region are due to SSS bending and the bands in the 400500 cm1
region are due to SS stretching (Ozin, 1969, Harvey and Butler, 1986, Parker et al., 2008, White, 2009).
The dissolution of chalcopyrite in sulphuric acid occurred with the formation of sulphur (area 2) and

showed no intense band at 471 cm-1 for areas (3) and (4), but rather vibrations at 268 cm-1, indicating the

IP

formation of covellite.

SC
R

Sulphur identified as a corrosion product on the chalcopyrite surface in the sulphuric acid may be
responsible for the higher passivation current density observed in this medium relative to the other acidic
media. The sulphur deposit reduces the anode area and the current density increases. The formation of

NU

elemental sulphur on the surface of chalcopyrite may have also contributed to the increase in
transpassivation potential.

MA

Elemental sulphur has been detected as a product of dissolution of chalcopyrite in acid media by means of
different techniques (White, 2009, Jones and Peters, 1976, Linge, 1976, Biegler and Swift, 1979, Muoz
et al., 1979, Majima et al., 1985). Its formation has been reported in a wide potential range of 0.5 to 1.5 V

vs. pH (Majuste et al., 2012). Hiroyoshi et al. (2004) using EIS to study the dissolution of chalcopyrite in

TE

the sulphuric acid reported the growth of a passive layer of elemental sulphur of high resistance on the
surface of chalcopyrite in sulphuric acid and in the absence of ferrous and cupric ions. Hiroyoshi et al.

CE
P

(2004) reported that in the passive region the Cu2S is not formed and only the passive elemental sulphur
layer growths according to the direct oxidation of chalcopyrite.
The covellite identified on the surface of the chalcopyrite after polarization tests in hydrochloric and nitric

AC

acids is a semiconductor (different to sulphur) allowing the transport of ions and electrons through the
corrosion product film. Warren et al. (1982) conducted a study of the anodic dissolution of chalcopyrite in
sulphuric acids based on thermodynamic analysis and proposed the formation of two intermediates, S1
and S2. The S1 product was responsible for the passivation of the sample potential near 0.5 V (SHE), and
has been proposed as bornite (Cu5FeS4). At higher potential than 0.5V (SHE), bornite (S1) oxidizes to
covellite, S2, (CuS) and subsequently to elemental sulphur (S).

4- Conclusions
The results of potentiodynamic anodic polarization show that the chalcopyrite dissolution is favored in
the hydrochloric acid. For leaching in the nitric acid, a concentration polarization was identified in the
cathodic branch of the polarization curve. The highest Warburg impedance was obtained for the

ACCEPTED MANUSCRIPT
dissolution of chalcopyrite in the sulphuric acid. The EIS results at low frequency region suggest that the
dissolution of chalcopyrite is controlled by the diffusion through the product layer in sulphuric and
hydrochloric acids.

IP

The characterization study of the electrodes by SEM / EDS and Raman spectroscopy revealed that the
chalcopyrite electrode presents cavities, faults, fractures and impurities on its surface. The use of Raman

SC
R

micro-spectroscopy allowed the identification of two corrosion products formed after polarization up to
1.3 V (Ag/AgCl), sulphur formed in the sulphuric acid, and covellite in the nitric and hydrochloric acids.

NU

Acknowledgement

The authors are grateful to the Brazilian government agencies: National Council for Scientific and

MA

Technological Development, Research Support Foundation of Minas Gerais State, Coordination of


Improvement of Higher Education Personnel. The authors thank Dr. Virginia Simpatico Teixeira
Ciminelli and Dr. Daniel Majuste, from the Federal University of Minas Gerais, to their valuable

contribution.

TE

References

Abdollahi, H., Shafaei, S.Z., Noaparast, M., Manafi, Z., Niemel, S.I., Tuovinen, O.H., 2014. Mesophilic

CE
P

and thermophilic bioleaching of copper from a chalcopyrite-containing molybdenite concentrate. Int. J.


Miner. Process. 128, 25-32.

Anderson, C., Fayram, T., Twidwell, L., 2011. An update on the NSC pressure oxidation of combined

AC

copper and molybdenum concentrates. 6th International Seminar on Copper Hydrometallurgy, 1-14.
Biegler, T., Swift, D.A., 1979. Anodic electrochemistry of chalcopyrite. J. Applied Electrochem 9, 545554.

Boukamp, B.A., 1985. Computer Aided Acquisition of Corrosion Data. Electrochem. Soc. 85-3, 146.
Crdoba, E.M., Muoz, J.A., Blzquez, M.L., Gonzlez, F., Ballester, A., 2008. Leaching of chalcopyrite
with ferric ion. Part I: General aspects. Hydrometallurgy 93, 81-87.
Dakubo, F., Baygents, J.C., Farrell, J., 2012. Peroxodisulphate assisted leaching of chalcopyrite.
Hydrometallurgy 121-124, 68-73.
Dreisinger, D., Abed, N., 2002. A fundamental study of the reductive leaching of chalcopyrite using
metallic iron part I: kinetic analysis. Hydrometallurgy 66, 37-57.

ACCEPTED MANUSCRIPT
Dreisinger, D., 2006. Copper leaching from primary sulphides: Options for biological and chemical
extraction of copper. Hydrometallurgy 83, 10-20.
Dutrizac, J.E., 1992. The leaching of sulphide minerals in chloride media. Hydrometallurgy 29, 1-45.

Feng, S., Yang, H., Zhan, X., Wang, W., 2014. Novel integration strategy for enhancing chalcopyrite

IP

bioleaching by Acidithiobacillus sp. in a 7-L fermenter. Bioresour. Technol. 161, 371-378.

SC
R

Ghahremaninezhad, A., Asselin, E, Dixon, D. G., 2010. Electrochemical evaluation of the surface of
chalcopyrite during dissolution in sulphuric acid solution. Electrochim. Acta 55, 50415056.
Ghahremaninezhad, A., Asselin, E., Dixon, D.G., 2012. Kinetics of the ferricferrous couple on

NU

anodically passivated chalcopyrite (CuFeS2) electrodes. Hydrometallurgy 125-126, 42-49.


Habashi, F., 2009. Copper Metallurgy: Past, Present, and Future, first ed. Antofagasta, Santiago.

MA

Harvey, P.D., Butler, I.S., 1986. Raman spectra of orthorhombic sulfur at 40 K. J. Raman
Spectrosc. 17, 329334.

Hiroyoshi, N., Kuroiwa, S., Miki, H., Tsunekawa, M., Hirajima, T., 2004. Synergistic effect of cupric

and ferrous ions on active-passive behavior in anodic dissolution of chalcopyrite in sulphuric acid

TE

solutions. Hydrometallurgy 74, 103-116.

Jones, D.L., Peters, E., 1976. The leaching of chalcopyrite with ferric sulphate and ferric chloride. In:

CE
P

Yannopoulos JC, Agarwal JC (eds.), International Symposium on Copper Extraction and Refining. The
Metallurgical Society of AIME, New York, United States of America. 2, 633-653.
Khoshkhoo, M., Dopson, M., Shchukarev, A., Sandstrm, A., 2014. Electrochemical simulation of redox

14.

AC

potential development in bioleaching of a pyritic chalcopyrite concentrate. Hydrometallurgy 144145, 7-

Li, Y., Kawashima, N., Li, J., Chandra, A.P., Gerson, A.R., 2013. A review of the structure, and
fundamental mechanisms and kinetics of the leaching of chalcopyrite. Adv. Colloid Interface Sci. 197, 132.
Li, Y., Qian, G., Li, J., Gerson, A.R., 2015. Kinetics and roles of solution and surface species of
chalcopyrite dissolution at 650 mV. Geochimica et Cosmochimica Acta. 161, 188-202.
Linge, H.G. 1976. A study of chalcopyrite dissolution in acidic ferric nitrate by potentiometric titration.
Hydrometallurgy 2, 51-64.
Lu, J., Dreisinger, D., 2013. Copper chloride leaching from chalcopyrite and bornite concentrates
containing high levels of impurities and minor elements. Hydrometallurgy 138: 40-47.

ACCEPTED MANUSCRIPT
Majima, H., Awakura, Y., Hirato, T., Tanaka, T., 1985. The leaching of chalcopyrite in ferric chloride
and ferric sulphate solutions. Can Metall Q 24, 283-291.
Majuste, D., Ciminelli, V.S.T., Osseo-Asare, K., Dantas, M.S.S., Magalhes-Paniago, R., 2012.

Electrochemical dissolution of chalcopyrite: Detection of bornite by synchrotron small angle X-ray

IP

diffraction and its correlation with the hindered dissolution process. Hydrometallurgy 111112, 114-123.

XRD

in

hydrometallurgy:

Literature

review and

Hydrometallurgy 131132, 54-66.

SC
R

Majuste, D., Ciminelli, V.S.T., Eng, P.J., Osseo-Asare, K., 2013. Applications of in situ synchrotron
investigation of

chalcopyrite

dissolution.

NU

McCafferty, E., 2010. Introduction to corrosion science. Springer: New York.


Mernagh, T.P., Trudu, A.G., 1993. A laser Raman microprobe study of some geologically important

MA

sulphide minerals. Chem Geol 103, 113-127.

Muoz, P.B., Miller, J.D., Wadsworth, M.E., 1979. Reaction Mechanism for the Acid Ferric Sulphate

Leaching of Chalcopyrite. Metall Trans B 10B, 149-158.

TE

Nicol, M., Miki, H., Velsquez-Yvenes, L., 2010. The dissolution of chalcopyrite in chloride solutions:
Part 3. Mechanisms. Hydrometallurgy 103, 86-95.

CE
P

Ozin, G.A., 1969. The single-crystal Raman spectrum of rhombic sulphur. J. Chem. Soc., A 116118.
Parker, G.K., Woods, R., Hope, G.A., 2008. Raman investigation of chalcopyrite oxidation. Colloids
Surf A: Physicochem Eng Aspects, 318, 160-168.

AC

Prasad, S., Pandey, B.D., 1998. Alternative processes for treatment of chalcopyrite a review. Miner.
Eng. 11, 763-781.

Qian, G., Li, Y., Gerson, A.R., 2015. Applications of surface analytical techniques in Earth Sciences.
Surface Science Reports 70 (1), 86-133.
Rammelt, U., Reinhard, G., 1990. On the applicability of a constant phase element (CPE) to the
estimation of roughness of solid metal electrodes. Electrochim. Acta 35, 1045-1049.
RRUFF Project, 2013. Integrated data base of Raman spectra, X-ray diffraction and chemistry data for
minerals. RRUFF ID: Chalcopyrite (R050559), aschamalmite (R060690), galenobismutite (R060988).
Available at http://rruff.info (January 2013).
Smith, Y.R., Sohn, H.Y., 2014. Application of additive-reaction-times law to the mixed-control kinetics
of oxygen leaching of chalcopyrite. Hydrometallurgy 146, 164-168.

ACCEPTED MANUSCRIPT
Sols-Marcal, O.J., Lapidus, G.T., 2014. Chalcopyrite leaching in alcoholic acid media. Hydrometallurgy
147148, 54-58.
Venkatachalam, S., 1991. Treatment of chalcopyrite concentrates by hydrometallurgical techniques.

Miner. Eng. 4, 1115-1126.

IP

Viramontes-Gamboa, G., Rivera-Vasquez, B.F., Dixon, D.G., 2007. Comparative Study Between

SC
R

Electrochemical and Leaching Responses. J. Electrochem. Soc. 154 C299-C311.

Wang, Y., Zeng, W., Chen, Z., Su, L., Zhang, L., Wan, L., Qiu, G., Chen, X., Zhou, H., 2014a.
Bioleaching of chalcopyrite by a moderately thermophilic culture at different conditions and community

NU

dynamics of planktonic and attached populations. Hydrometallurgy 147148, 13-19.


Wang, Y., Su, L., Zeng, W., Wan, L., Chen, Z., Zhang, L., Qiu, G., Chen, X., Zhou, H., 2014b. Effect of

MA

pulp density on planktonic and attached community dynamics during bioleaching of chalcopyrite by a
moderately thermophilic microbial culture under uncontrolled conditions. Miner. Eng. 61, 66-72.
Warren, G.W., Wadsworth, M.E., El-Raghy, S.M., 1982. Passive and Transpassive Anodic Behavior of

Chalcopyrite in Acid Solutions. Metall. Trans. B 13 B, 571-579.

TE

Watling, H.R., 2014. Chalcopyrite hydrometallurgy at atmospheric pressure: 2. Review of acidic chloride
process options. Hydrometallurgy 146, 96-110.

CE
P

White, S.N., 2009. Laser Raman spectroscopy as a technique for identification of seafloor hydrothermal
and cold seep minerals. Chem. Geol. 259, 240-252.
Yu, J., Yang, H., Fan, Y., 2011. Effect of potential on characteristics of surface film on natural

AC

chalcopyrite. Trans.Nonferrous Met. Soc. China 21, 1880-1886.


Zhang, L., Wu, J., Wang, Y., Wan, L., Mao, F., Zhang, W., Chen, X., Zhou, H., 2014. Influence of
bioaugmentation with Ferroplasma thermophilum on chalcopyrite bioleaching and microbial community
structure. Hydrometallurgy 146, 15-23.
Zheng, Z.J., Gao, Y., Gui, Y., Zhu, M., 2014. Studying the fine microstructure of the passive film on
nanocrystalline 304 stainless steel by EIs, XPS and AFM, J.Solid State Electrochem., 2014, 18, 22012210.

ACCEPTED MANUSCRIPT
Figure Captions
Figure 1 Anodic polarization curves of chalcopyrite from 0.1 V(Ag/AgCl) to 1.3 V(Ag/AgCl) in
0.1mol/L aqueous solution of HCl, HNO3 and H2SO4 at room temperature and scan rate of 50 mV.s-1.

IP

Figure 2 Open circuit potential in function of time in 0.1mol/L aqueous solutions of HCl, HNO3 and
H2SO4 at room temperature.

SC
R

Figure 3 Nyquist (a) and Bode (b) diagrams of chalcopyrite in 0.1mol/L aqueous solutions of HCl,
HNO3 and H2SO4 at room temperature, potential amplitude of 10mV and frequency range were of 105 10-2 Hz.

NU

Figure 4 Equivalent circuit fitted to the EIS data: (a) for the solutions of hydrochloric and sulphuric
acids and (b) for the nitric acid.

MA

Figure 5 SEM micrograph of chalcopyrite (a) after polishing (b) after polarization in solution of
0.1mol/L HCl at 25 C.

Figure 6 SEM micrograph of chalcopyrite (a) after polishing (b) after polarization in solution of

TE

0.1mol/L HNO3 at 25 C.

Figure 7 SEM micrograph of chalcopyrite (a) after polishing (b) after polarization in solution of

CE
P

0.1mol/L H2SO4 at 25 C.

Figure 8 Optical micrograph and Raman spectrum of chalcopyrite electrode after polishing and before
polarization tests - region 1 (a).

AC

Figure 9 - Optical micrograph and Raman spectra of chalcopyrite electrode after polishing and before
polarization tests - region 1 (b).
Figure 10 Optical micrograph of chalcopyrite electrode after polarization in HCl solution and Raman
spectra of areas 2, 3 and 4.
Figure 11 Optical micrograph of chalcopyrite electrode after polarization in HNO 3 solution and Raman
spectra of areas 2, 3 and 4.
Figure 12 - Optical micrograph of chalcopyrite electrode after polarization in H 2SO4 solution and Raman
spectra of areas 2, 3 and 4.

MA

NU

SC
R

IP

ACCEPTED MANUSCRIPT

AC

CE
P

TE

Figure 1

TE

MA

NU

SC
R

IP

ACCEPTED MANUSCRIPT

AC

CE
P

Figure 2

AC

Figure 3

CE
P

TE

MA

NU

SC
R

IP

ACCEPTED MANUSCRIPT

ACCEPTED MANUSCRIPT

AC

CE
P

TE

MA

NU

SC
R

IP

Figure 4a

IP

ACCEPTED MANUSCRIPT

AC

CE
P

TE

MA

NU

SC
R

Figure 4b

SC
R

IP

ACCEPTED MANUSCRIPT

AC

CE
P

TE

MA

NU

Figure 5

SC
R

IP

ACCEPTED MANUSCRIPT

AC

CE
P

TE

MA

NU

Figure 6

SC
R

IP

ACCEPTED MANUSCRIPT

AC

CE
P

TE

MA

NU

Figure 7

NU

SC
R

IP

ACCEPTED MANUSCRIPT

AC

CE
P

TE

MA

Figure 8

SC
R

IP

ACCEPTED MANUSCRIPT

AC

CE
P

TE

MA

NU

Figure 9

NU

SC
R

IP

ACCEPTED MANUSCRIPT

AC

CE
P

TE

MA

Figure 10

NU

SC
R

IP

ACCEPTED MANUSCRIPT

AC

CE
P

TE

MA

Figure 11

NU

SC
R

IP

ACCEPTED MANUSCRIPT

AC

CE
P

TE

MA

Figure 12

ACCEPTED MANUSCRIPT
Table 1 Electrochemical parameters obtaining using anodic polarization technique
Solution icrit.

Epass.

Etransp.

(Etransp.-

(A/cm2)

(mV)

(mV)

Epass.)
(mV)

SC
R

IP

(A/cm2)

ipass.

1.6x10-51.2x10-6

1.4x10-51.2x10-6

35721

38827

31

H2SO4

6.4x10-43.8x10-5

1.2x10-47.2x10-6

39728

75067

353

HNO3

6.3x10-54.4x10-6

1.79x10-58.9x10-7

38935

52531

136

AC

CE
P

TE

MA

NU

HCl

ACCEPTED MANUSCRIPT
Table 2 Fe and Cu concentrations after anodic polarization tests
Fe

Fe

Cu

Cu

solution

concentration

concentration concentration

(mg/L)

(mol/L)

(mol/L)

SC
R

IP

(mg/L)

Cu/Fe

concentration ratio

Acid

(molar
ratio)

0.200.01

3.54x10-6

0.280.01

4.39x10-6

1.24

HNO3

0.220.01

3.92x10-6

0.140.01

2.17x10-6

0.55

H2SO4

0.180.01

3.21x10-6

0.320.02

5.08x10-6

1.58

AC

CE
P

TE

MA

NU

HCl

ACCEPTED MANUSCRIPT
Table 3 Open circuit potential and the electrochemical parameters obtaining from EIS

R2(cm2)

IP

7.70.6

SC
R

R1(cm2)

HNO3
209 (4)

H2SO4
220 (10)
5.20.3

6000759

12518751

4000240

R3(cm2)

12000480

1672904942

123411309

Ws-P

0.4034

CPE1, T(Fsncm-2)

2.3x10-5

1.7x10-5

1.5x10-5

CPE2, T(Fsncm-2)

1.1x10-5

2.4x10-5

4.1x10-6

Chi-squared

2.3x10-4

1.8x10-4

2.8x10-4

MA

TE
CE
P
AC

NU

8.50.5

Emixed(mV) Ag/AgCl

HCl
202 (27)

measurements

0.4513

AC

CE
P

Graphical abstract

TE

MA

NU

SC
R

IP

ACCEPTED MANUSCRIPT

ACCEPTED MANUSCRIPT
Highlights
The acid dissolution of chalcopyrite is studied using electrochemical impedance

IP

spectroscopy. Potentiodynamic polarization of chalcopyrite in acid media was


performed. The hydrochloric acid is a more efficient leaching agent than nitric and

SC
R

sulphuric acids. Dissolution of chalcopyrite is controlled by diffusion. Lead


and bismuth sulphides were identified as impurities on the mineral surface. Sulphur

AC

CE
P

TE

MA

NU

and covellite were identified as corrosion products on the mineral surface.

Potrebbero piacerti anche