You are on page 1of 9

Acta Metallurgica Slovaca, 10, 2004, Special Issue 2 (174 - 182) 174

LEACHING OF CHALCOPYRITE IN CUPRIC CHLORIDE SOLUTION

Lundström M.1, Aromaa J.1, Forsén O.1, Hyvärinen O.2 and Barker M. H.2
1
Helsinki University of Technology, Laboratory of Corrosion and Material Chemistry,
P.O. Box 6200, 02015 HUT, Finland, e-mail: Mari.Lundstrom@hut.fi
2
Outokumpu Research Oy, Kuparitie 10, P.O. Box 60, 28101 Pori, Finland

LÚHOVANIE CHALKOPYRITU V ROZTOKU CHLORIDU MEĎNATÉHO

Lundström M.1, Aromaa J.1, Forsén O.1, Hyvärinen O.2, Barker M. H.2
1
Helsinki University of Technology, Laboratory of Corrosion and Material Chemistry,
P.O. Box 6200, 02015 HUT, Finland, e-mail: Mari.Lundstrom@hut.fi
2
Outokumpu Research Oy, Kuparitie 10, P.O. Box 60, 28101 Pori, Finland

Abstrakt
Outokumpu HydroCopperTM technológia bola vyvinutá Outokumpu Research. Je to
nová metóda výroby medi založená na účinných lúhovacích vlastnostiach roztoku chloridu
meďnatého. Lúhovanie sa realizuje pri atmosferickom tlaku, blízko bodu varu vody v reaktore
s miešaním roztoku, ktorý je udržiavaný na pH 1.5-2.5.
V práci bolo realizované elektrochemické štúdium chovania sa chalkopyritu,
najbežnejšieho minerálu medi, počas jeho rozpúšťania v roztoku koncentrovaného chloridu
sodného (250 g/l) s premenlivou koncentráciou meďnatých iónov v rozsahu 0.09-26.6 g/l.
Taktiež bol meraný vplyv pH (1-3) a teploty (70-90 °C).
Pre určenie mechanizmu rozpúšťania chalkopyritu boli elektródy analyzované
pomocou elektrónovej skenovacej mikroskópie. Pre určenie rýchlosť kontrolujúceho kroku
lúhovania bola použitá metóda rotujúceho disku. So zvyšujúcou sa teplotou a koncentráciou
meďnatých iónov pre [Cu2+] > 9 g/l bolo pozorované proporcionálne zvýšenie reakčnej
rýchlosti. Nižšia koncentrácia meďnatých iónov, t.j. [Cu2+] < 9 g/l, nevplývala na rýchlosť
lúhovania. Výsledky nasvedčujú tomu, že pri vyššom pH sa na povrchu lúhovaného
chalkopyritu hromadia oxidy, hydroxidy a chloridy železa. Pri nižších hodnotách pH sa
chalkopyrit pasivuje a na jeho povrchu bola detekovaná vrstva obohatená o síru.

Abstract
Outokumpu HydroCopperTM technology was developed by Outokumpu Research. It
is a novel copper production method based on the efficient leaching properties of cupric
chloride solutions. The leaching is operated at atmospheric pressure, near the boiling point of
water in a stirred reactor and the pH is kept between 1.5-2.5.
In the present work, an electrochemical study of the dissolution behaviour of
chalcopyrite, the most common copper mineral, was made in a concentrated sodium chloride
solution (250 g/l) with variable cupric ion concentrations in the range 0.09–26.6 g/l. The
effects of pH (1-3) and temperature (70–90 ºC) were also measured.
In order to investigate the mechanism of the chalcopyrite dissolution, the electrodes
were analysed by scanning electron microscopy. The rotating disk electrode technique was
used to define the rate controlling step in the leaching process. The increase in the reaction
rate was observed to be proportional to the increase in the temperature and to the cupric ion
concentration with [Cu2+] > 9 g/l. With [Cu2+] < 9 g/l the cupric ion concentration did not
effect the leaching rate. The results suggested that iron oxides, hydroxides and chlorides
Acta Metallurgica Slovaca, 10, 2004, Special Issue 2 (174 - 182) 175

gathered on the chalcopyrite surface at higher pHs. At lower pHs chalcopyrite passivated and
a sulfur rich layer was detected on the surface.

Key words: Chalcopyrite, leaching of chalcopyrite, cupric chloride, sodium chloride,


HydroCopperTM

1. Introduction
There are a great number of research articles on the leaching of sulphide minerals in
chloride media. In highly concentrated chloride solution copper ions readily form cuprous
complexes, such as [CuCl3]2-, [Cu2Cl4]2-, [Cu3Cl6]3- and cupric complexes, such as [CuCl]+,
[CuCl2]0, [CuCl3]-, [CuCl4]2- [1-3]. The utilization of chloride solutions in chalcopyrite
leaching is advantageous due to the aggressive nature of the leaching and to the stability of
cuprous ions due to the formation of the aforementioned chloro-complexes. With pHs > 3,
cupric hydroxyl chlorides become important [4]. In general, the leaching of chalcopyrite is
more effective in chloride solutions with cupric ions as the oxidant than in sulfate solutions
with ferric ions as the oxidant. This is possibly due to kinetic rather than thermodynamic
considerations e.g. the higher rates of electron transfer in chloride solutions than in sulphate
solutions, as the chalcopyrite surface passivates more readily in the presence of sulfate ions.
[5] The formation of passivating reaction product films has been suggested several times in
sulfate solutions [6, 7]. Munoz et al. [8] concluded that the rate limiting step in the leaching of
chalcopyrite with ferric sulfate was the transport process through the sulfur product layer.
Similar behaviour has also been reported in ferric chloride leaching [9].
Recently, Outokumpu has developed a novel leaching process, HydroCopperTM. In
this process [10,11] as well as in the modified Ecuprex® [12] process chalcopyrite is reported
to dissolve according to reaction (1). [13]

CuFeS2(s) + 3Cu2+(aq) → 4Cu+(aq) + Fe2+(aq) + 2S0(s) (1)

Wilson et al. [14] suggested similar leaching reactions in cupric chloride solutions,
however, via the formation of cuprous chloride complexes (2).

CuFeS2(s) + 3[CuCl]+(aq) + 11Cl-(aq) = 4[CuCl3]2-(aq) + FeCl2(aq) + 2S0(s) (2)

According to Wilson et al. [14] the ratio of cuprous to cupric ions must be less than
1.9 for reaction (2) to be thermodynamically favourable. They also observed that over the
concentration range [Cu2+] = 0.79 - 1.46 M and [Cl-] = 2.82 - 6.21 M at 90 ºC, neither cupric
nor chloride ion concentration influences the rate of chalcopyrite dissolution. In
HydroCopperTM the leaching of chalcopyrite is done in a cupric chloride solution at
atmospheric pressure and at a temperature of 80-100 ˚C. The pH is kept in between 1.5 and
2.5. [10,11] The aim of this work is to build a process window of the behaviour of
chalcopyrite in a highly concentrated chloride solution. The effect of temperature, pH and
cupric ion concentration on the dissolution of chalcopyrite is studied.

2. Thermodynamic
The standard free energy of formation for chalcopyrite is –242.70 kJ/mol. The
formation of solvated cupric and cuprous ions is not thermodynamically favourable under
standard conditions; the standard free energy of formation, ∆G0, for Cu+ is 50.20 kJ/mol and
Acta Metallurgica Slovaca, 10, 2004, Special Issue 2 (174 - 182) 176

for Cu2+ is 64.96 kJ/mol. However, the formation of chloro-complexes is thermodynamically


favourable. [14] When a system is not in a standard state, the equilibrium potential of the
reaction can be calculated with the Nernst equation. To draw Fig. 1, the E0 values were
calculated for all equations with HSC 5.11 software by Outokumpu Research. These values
were imported in to Microsoft Excel and the effect of chloride ion concentration was taken in
to account by using the Nernst equation.
0.7

0.6
+
C uC l

0.5
2+
Cu
0.4
2-
C uC l2
- C uC l3
Eh

0.3
Cu

0.2

0.1

-0.1
-2 -1 .5 -1 -0 .5 0 0 .5 1
log (cC l-)

Fig. 1. E - log c(Cl-) diagram showing the stability of metallic copper, cupric ions and cupric
and cuprous complexes at T = 90 °C. [Cu2+] = 1 M (solid line) and [Cu2+] =0.1 M
(dashed line).

Fig. 1 shows the effect of cupric ion concentration (0.1 M and 1 M) on the stability of
copper, cupric ions and cupric and cuprous complexes at 90 °C. Increasing the copper
concentration decreases the stability of cupric ions and makes the formation of [CuCl]+ and
[CuCl2]- more likely. Also the effect of temperature on the stabilities was studied. The stable
area of cupric ions decreases as the temperature increases. The [CuCl2]- complex is more
stable at 90 °C than at 25 °C. Otherwise the changes in the stabilities within the temperature
range 25–90 °C are small. According to the E-c(Cl-) diagrams, under HydroCopperTM
conditions the dominant complex is likely to be [CuCl]+. That suggests that the leaching of
chalcopyrite would more likely occur via the formation of cuprous chloride complexes similar
to reaction (2) than according to reaction (1).

3. Materials and procedures


Electrochemical methods such as anodic polarization, potentiostatic measurements
and cyclic voltammetry were used in this study. A rotating disk electrode (RDE) was used to
determine the leaching rate controlling step. Scanning electron microscopy / energy dispersive
(X-ray) spectrocopy (SEM/EDS) was applied to study the composition of reaction product
layers formed on the chalcopyrite.
The concentration of NaCl in the solution was 250 g/l (4.3 M). The cupric ion
concentrations used were 0.09, 0.9, 4.5, 9.0, 17.9 and 26.6 g/l. The temperature was varied
between 70 and 90 ˚C and the pH was adjusted in the range of 1 to 3. The upper pH limit is
defined by the formation of copper oxychlorides.
The working electrodes were made of chalcopyrite from the Pyhäsalmi mine in
Finland. The elemental composition of the electrodes was analysed several times with
Acta Metallurgica Slovaca, 10, 2004, Special Issue 2 (174 - 182) 177

SEM/EDS. The composition varied, but the average value was Cu 30.0%, Fe 32.5%, S 36.3%,
Si 0.5%, Al 0.4% and Mg 0.3% in percentage by weight. The analysed values were near to the
theoretical composition of chalcopyrite (Cu 34.6 %, Fe 30.4 % and S 34.9 %). The electrodes
were polished between every measurement with rotating and wetted grade 800 waterproof
abrasive paper. After polishing the electrodes were rinsed first with de-ionised water, then
ethanol and then dried.
A standard three-electrode electrochemical cell was employed for the electrochemical
measurements. The vessel was equipped with a thermostated water jacket. In all measurements
(except RDE) the cell was stirred with a magnetic stirrer at 500 r.p.m. No purging of gases
was done. The counter electrode was a platinum sheet, the reference electrode was Ag/AgCl
placed in a sintered glass tube containing gel of agar powder, potassium chloride and distilled
water. The reference electrode junction was positioned in an external beaker and connected to
the cell via a Luggin capillary. The measurements were carried out using two electrochemical
workstations: (i) a PAR 273 Potentiostat/Galvanostat controlled by EG&G PAR´s Model 352
Corrosion Analysis Software 1.00 and (ii) Potentiostat/Galvanostat 2000 working together
with a 5050 frequency response analyser (FRA) and 1731 Intelligent/Arbitrary function
synthesizer (both NF Corporation, Japan) controlled by in-house software.
The dissolution current densities were estimated from anodic polarization curves with
the Tafel method. The rotating disc technique was used to study the mass transport as the
Levich equation (3) predicts the variation in the transport-limited current as a function of the
electrode rotation rate

jlim = 0.62 zFD2/3ν-1/6ω1/2c (3)

where j is the current density (mA/cm2), F is Faraday’s constant (96,485 C/mol), D is the
diffusion coefficient (cm2/s), ν is the kinematic viscosity (cm2/s), ω is the angular frequency
(1/s (=2πf)) and c is the concentration (mol/cm3). In the Levich plot the transport limited
current as a function of the square root of rotation speed should yield a straight line. If the plot
gives a linear response, the system can be assumed to be diffusion controlled and the diffusion
coefficient can be calculated.

4. Results
4.1. The effect of [Cu2+]
With cupric ion concentration ≤ 9 g/l (0.14 M) at pH 2 and 85 °C the corrosion
potential was observed to follow the Nernst equation. The increase in the corrosion potential
was ca. 60 mV/decade of cupric ion concentration (Fig. 2). This is in good agreement with the
study done by Hirato et al. [15]. They found the effect of CuCl2 concentration (from 0.01 to 1
mol/dm3) on the mixed potential of chalcopyrite to be 66 mV/decade at 70 °C in 0.2mol/dm3
(7.3 g/l) HCl solution. At cupric ion concentrations > 9 g/l the behaviour changed and the
corrosion potential did not follow the Nernst equation.
The effect of copper concentration was studied with anodic polarisation
measurements. Increasing the cupric ion concentration, when [Cu2+] > 9 g/l, increased the
dissolution rate of the chalcopyrite for the temperature range studied (Fig. 3). When pH ≥ 2.5,
or the [Cu2+] ≤ 9 g/l, the cupric ion concentration did not seem to play an important role in
the dissolution process .
Acta Metallurgica Slovaca, 10, 2004, Special Issue 2 (174 - 182) 178

580

560 70°C
75°C
80°C

POTENTIAL (mV vs. Ag/AgCl)


540
85°C
90°C
520

500

480

460

440

0.1 1 10
CUPRIC ION CONCENTRATION (g/l)
Fig. 2. A plot of corrosion potential vs. cupric ion concentration as a function of temperature.
[NaCl] = 250 g/l and pH = 2.

1.5

Cu(II) 26.6 g/l


Cu(II) 17.9 g/l
Cu(II) 9.0 g/l
CURRENT DENSITY (mA/cm )

Cu(II) 4.5 g/l


2

1.0
Cu(II) 0.9 g/l
Cu(II) 0.09 g/l

0.5

0.0
400 500 600 700 800 900 1000

POTENTIAL (mV vs. Ag/AgCl)


Fig. 3. Anodic polarisation curves run with cupric ion concentration 0.09, 0.9, 4.5, 9.0, 17.9
and 26.6 g/l. [NaCl] = 250 g/l, T = 90 ºC, pH = 2 and scan rate = 0.33 mV/s.

The dissolution rates calculated by the Tafel method also changed remarkably when
the cupric ion concentration was increased from 9 to 17.9 g/l. With cupric concentrations
above 9 g/l the dissolution increased to over 5 µm/h, this increased with increasing cupric
concentration. With cupric concentrations below 9 g/l the dissolution rate was below 2 µm/h.
The reaction rate controlling mechanism appears to change with cupric ion concentration of 9
g/l. This is discussed in more detail in section 4.4.

4.2. The effect of pH on the chalcopyrite dissolution


The effect of pH on the corrosion potential and the anodic polarisation curves was
remarkable. A low pH, in the pH range 1 to 2.5, was observed to decrease the corrosion
Acta Metallurgica Slovaca, 10, 2004, Special Issue 2 (174 - 182) 179

potential by ca. 20 mV/pH unit. Since protons are not involved in dissolution reactions (1) or
(2), a possible explanation is that a reaction product layer forms, which then reacts with
protons.
6

pH 1
5 pH 1.5
pH 2

CURRENT DENSITY (mA/cm )


2
pH 2.5
4 pH 3

0
500 600 700 800 900 1000
POTENTIAL (mV vs. Ag/AgCl)

Fig. 4. Anodic polarisation curves at pHs 1.0, 1.5, 2.0, 2.5 and 3.0, [NaCl] = 250 g/l, [Cu2+] =
26.6 g/l, T = 70 ºC and scan rate = 0.33 mV/s.

Anodic polarization measurements were carried out in the pH range 1 to 3. It was


observed that the current densities at pH 2.5 and 3 were remarkably greater than those at lower
pHs, Figure 4. The colour of the electrode surface after the experiments depended on the pH:
at 1 and 1.5 the surface was golden, at 2 and 2.5 black, gray and brown and at pH 3 rust
coloured.
Potentiostatic measurements (not shown) indicated that the reaction mechanism
changed between pH 2 and pH 2.25. At pH < 2.25 the current density decreases very quickly,
which suggests the formation of a passive layer on the surface. Cyclic voltammograms done at
pH 2 supports this as no peaks were seen. Quantitative calculations from the SEM analysis on
the electrode surface after the measurement at pH 1.5 showed only copper, iron and an excess
of sulfur. Thus the passive layer is suggested to be sulfur rich or possibly pure sulfur. It was
also observed that at pH ≥ 2.25 there is a period of constant current until t = 1100s, after which
the current density starts to decrease.
Voltammograms at pH 3 showed two peaks at 740 and 870 mV vs. Ag/AgCl, which
are most likely due to the formation of the rust-coloured layer. The current density decreased
with each subsequent sweep and the reaction rate finally became constant. This indicates that a
reaction product layer forms and the reaction taking place through that layer reaches steady
state.
The cross cut analysis by SEM/EDS, of an electrode after anodic polarization at pH 3
showed the formation of a two-phase layer consisting mainly of iron and oxygen (in the ratio
1:2 to 1:3). Low concentrations of chloride and sulfur were also present, but almost no copper
or sodium were detected. The colour of the electrode surface after anodic polarization at pH 3
was red with cupric concentrations of 17.9 g/l or greater, and more yellow with lower cupric
concentrations. This suggests that at pH 3, with higher cupric ion concentrations, a hematite-
like layer is formed on the surface, and with lower cupric ion concentrations a hydrated iron
oxide or goethite type layer is formed. [16]

4.3. Temperature Effect


In the temperature range 70 to 90 °C at pH 2 an increase in temperature gave a
corresponding increase in the corrosion potential for all cupric ion concentrations. From the
Acta Metallurgica Slovaca, 10, 2004, Special Issue 2 (174 - 182) 180

anodic polarisation measurements it was observed that the increase in temperature was
accompanied by an increase in the corrosion current density with all cupric ion concentrations
at pH 2. The current densities doubled when the temperature was increased from 70 °C to 90
°C.

4.4. Reaction controlling mechanism


The polarization curves both with standard and rotating disk electrodes were used to
clarify the mechanism of the rate-controlling step. The activation energy, Ea, was calculated at
600, 700 and 800 mV vs. Ag/AgCl. Levich plots were made at 700, 800 and 900 mV vs.
Ag/AgCl. From the anodic polarization curves when the cupric concentration was 4.5 or 9 g/l
(pH = 2, T = 85 °C) then Ea was calculated to be 40-45 kJ/mol. With higher copper
concentrations (17.9 and 26.6 g/l) Ea was < 35 kJ/mol. The Levich plot gave a linear response
with a cupric ion concentrations of 9 g/l and 17.9 g/l (pH 2, T = 85 °C), Fig. 5. With a cupric
concentration of 17.9 g/l the diffusion coefficient was calculated with Levich equation to be
4.4 × 10-6 cm2/s, when the cupric concentration was 9.0 g/l the diffusion coefficient was ca.
1/10th of that value. Again, this supports the idea that a reaction product layer is formed on the
chalcopyrite surface and the reactions on that layer are diffusion controlled. With cupric ion
concentrations 0.9 and 4.5 g/l, the Levich plot was non-linear.

1.2
at 700 mV vs. Ag/AgCl
1.1 y = 0.044x + 0.18
LIMITING CURRENT DENSITY (mA/cm )
2

1.0 at 800 mV vs. Ag/AgCl


y = 0.054x + 0.12
0.9 at 900 mV vs. Ag/AgCl
0.8 y = 0.056x + 0.11

0.7

0.6

0.5

0.4

0.3

0.2

0.1

0.0
0 2 4 6 8 10 12 14 16 18
SQUARE ROOT OF ANGULAR SPEED (1/s)

Fig. 5. Levich plot with cupric ion concentration 17.9 g/l. Limiting current density taken at
700, 800 and 900 mV vs. Ag/AgCl

We propose that with cupric concentrations greater than 9g/l, at pH = 2 and T = 85


°C, the rate controlling factor is diffusion in the solution. With a cupric concentration of 9 g/l
the reaction is controlled by diffusion through the reaction product layer. With lower cupric
ion concentrations the reaction is under mixed or chemical control.

5. Discussion and conclusions


For pHs in the range 1.0 to 2.5 and temperatures in the range 70 to 90 °C a critical
copper concentration of 9 g/l was found. Below that concentration limit the corrosion potential
followed the Nernst equation and the increase in the corrosion potential was ca. 60
mV/decade. It was also found that [Cu2+] ≤ 9 g/l did not have any effect on the reaction rate.
Cupric concentrations above 9 g/l increased the reaction rate. The same was calculated from
Acta Metallurgica Slovaca, 10, 2004, Special Issue 2 (174 - 182) 181

Tafel plots, i.e. that [Cu2+] ≤ 9 g/l did not affect the dissolution rate and concentrations greater
than 9 g/l increased the dissolution rate.
The reaction rate control appeared to change around [Cu2+] = 9 g/l. Above that the
reaction rate control is assumed to be diffusion in solution, the diffusion coefficient being
ca. 4.4 × 10-6 cm2/s. With [Cu2+] = 9 g/l the reaction is controlled by diffusion through the
reaction product layer. With lower cupric ion concentrations the reaction was under chemical
or mixed control. An increase in temperature between 70 and 90 °C (at pH 2) was observed to
increase the current densities and the dissolution rates for all cupric ion concentrations.
The results indicate that there is a change in the electrochemical behavior of the
system between pH 2 and 2.25 (with [Cu2+] > 9 g/l and T = 70-90 °C). Low dissolution rates
at more acidic pH values are due to the rapid formation of a sulfur-rich reaction product layer
that impedes the dissolution of chalcopyrite, supported by the detection of excess surface
concentrations of sulfur from the SEM/EDS analysis. The dissolution rates were higher at less
acidic pHs and this is most likely due to the slower formation of the porous iron rich reaction
product layer. The Fe:O ratio was calculated to be between 1:2 and 1:3.
Together with visual observations the presence of goethite and iron hydroxides is
suggested.

1.2 2+
[Cu ] > 9 g/l and pH > 2
CORROSION CURRENT DENSITY (mA/cm )
2

2+
[Cu ] > 9 g/l and pH ≤ 2
2+
1.0 [Cu ] ≤ 9 g/l and pH > 2
2+
[Cu ] ≤ 9 g/l and pH ≤ 2

0.8

0.6

0.4

0.2

0.0
440 460 480 500 520 540 560 580 600
POTENTIAL (mV vs. Ag/AgCl)

Fig. 6. A plot of corrosion current density vs. corrosion potential.(i) [Cu2+] > 9 g/l and pH >
2.25, (ii) [Cu2+] > 9 g/l and pH < 2.25, (iii) [Cu2+] ≤ 9 g/l and pH > 2.25 and (iv) [Cu2+]
≤ 9 g/l and pH < 2.25.

The value of jcorr was shown to depend not only on the corrosion potential, but it is
also strongly dependent on several solution parameters. The corrosion current density in Fig. 6
was determined from the anodic polarization curves by the Tafel method. The highest
calculated values of jcorr were ca. 1 mA/cm2. A process window, where the dissolution of
chalcopyrite is possible, can be formed. The pH should be > 2 or the formation of a sulfur-rich
product layer effectively prevents the leaching. The upper limit for pH is the precipitation of
copper oxychloride which begins around pH 3. Cupric ion concentration should be > 9 g/l.
The best leaching environment is thus one of high cupric ion concentration, high pH and high
temperature, as shown in group (i) of Fig. 6.
Acta Metallurgica Slovaca, 10, 2004, Special Issue 2 (174 - 182) 182

References
[1] Berger, J.M., Winand, R., Solubilities, densities and electrical conductivities of aqueous
copper(I) and copper (II) chlorides in solutions containing other chlorides such as iron,
zinc, sodium and hydrogen chlorides. Hydrometallurgy 12(1984), pp. 61-81.
[2] Fritz, J.J., Chloride Complexes of CuCl in Aqueous Solution. Journal of Physical
Chemistry 84(1980) 18, pp. 2241-2246.
[3] Fontana, A., van Muylder, J., Winand, R., Etude Spectrophotometrique De Solutions
Aqueuses Chlorurees De Chlorure Cuivreux, a Concentrations Elevees. Hydrometallurgy
11(1983), pp. 287-314.
[4] McDonald, G.W., Udovic, T.J., Dumesic, J.A., Langer, S.H., Equilibria associated with
cupric chloride leaching of chalcopyrite concentrate. Hydrometallurgy 13(1984), pp.
125-135.
[5] Muir, D.M., Basic Principles of Chloride Hydrometallurgy. Chloride Metallurgy
II(2002), pp. 759-778.
[6] Hackl, R.P., Dreisinger, D.B., Peters, E., King, J.A., Passivation of Chalcopyrite during
oxidative leaching in sulfate media. Hydrometallurgy 39(1995), pp. 25-48.
[7] Parker, A., Klauber, C., Kougianos, A., Watling, H.R., van Bronswijk, W., An X-ray
photoelectron spectroscopy study of the mechanism of oxidative dissolution of
chalcopyrite. Hydrometallurgy 71(2003), pp. 265-276.
[8] Munoz, P.B., Miller, J.D., Wadsworth, M.E., Reaction Mechanism for the Acid Ferric
Sulfate Leaching of Chalcopyrite. Metallurgical Transactions B, 10B(1979), pp. 149-
158.
[9] Roman, R.J., Benner, B.R., The dissolution of copper concentrates. Minerals Science
and Engineering 5(1973) 1, pp. 3-24.
[10] Hietala, K., Hyvärinen, O., HydroCopperTM – A New Technology for Copper
Production. In Alta 2003 Copper Conference. Perth, Australia, 22-23 May.
[11] Hyvärinen, O., Hämäläinen, M., Leimala, R., Outokumpu HydroCopperTM Process A
Novel Concept in Copper Production. In Chloride Metallurgy 2002, 32nd Annual
Hydrometallurgy Meeting. Peek, E., van Weert, G. (eds). Montreal, Qubec, Canada.
MetSoc. pp. 609-612.
[12] Olper, M., Maccagni, M., The modified Ecuprex process: A promising hydrometallurgy
approach for chalcopyrite-bearing copper concentrates. In Hydrometallurgy of Copper.
Riveros, P.A., Dixon, D.G., Dreisinger, D.B., Menacho, J.H. (eds.). Santiago, Chile,
2003. Met. Soc., pp. 319-334.
[13] Habashi, F., Chalcopyrite its Chemistry and Metallurgy. McGraw-Hill, Great Britain,
1978, pp. 1, 63, 78, 79.
[14] Wilson, J.P., Fisher, W.W., Cupric Chloride Leaching of Chalcopyrite. Journal of Metals
33(1981) 2, p. 52-57.
[15] Hirato, T., Majima, H., Awakura, Y., The Leaching of Chalcopyrite with Cupric
Chloride. Metallurgical Transactions B, 18B(1987), pp. 31-39.
[16] Lundström, M., Leaching of chalcopyrite in cupric chloride media. M. Sc. Thesis,
Helsinki University of Technology, Espoo, Finland. 2004. p. 96.

You might also like