Sie sind auf Seite 1von 8

Hydrometallurgy 85 (2007) 9 16

www.elsevier.com/locate/hydromet

Cathodic reactions of Cu 2+ in cupric chloride solution


Mari Lundstrm a,, Jari Aromaa a , Olof Forsn a , Olli Hyvrinen b , Michael H. Barker b
a

Helsinki University of Technology, Laboratory of Corrosion and Materials Chemistry, P.O. Box 6200, 02015 HUT, Finland
b
Outokumpu Research Oy, P.O. Box 60, 28101 Pori, Finland
Received 22 December 2005; received in revised form 12 June 2006; accepted 7 July 2006
Available online 1 September 2006

Abstract
In this study we demonstrate the kinetics of Cu2+ reduction in concentrated cupric chloride solutions. Experiments were carried
out near the boiling point of the solution ([NaCl] = 280 g/l and [Cu2+] = 140 g/l) at T = 90 C, atmospheric pressure, pH = 2.
Electrochemical methods such as cathodic polarization curves and cyclic voltammetry were used to investigate the cathodic
reactions of copper complexes. To identify the nature and the rate-controlling steps of the reactions, rotating disk electrode (RDE)
experiments were conducted. The chemical environment studied was similar to that of the Outokumpu HydroCopperTM process,
which uses a cupric chloride solution to leach copper from the mineral chalcopyrite.
The results suggest that the cathodic reactions are the reduction of [CuCl]+ to the complex [CuCl3]2, the reduction of [CuCl3]2
to solid copper and hydrogen evolution. The diffusion coefficient and the unit rate constants for the solution species were
calculated. The exchange current density and rate constant for electron transfer were also estimated. A simulation was made of the
cathodic polarization curve and it was in good agreement with the experimental data.
2006 Elsevier B.V. All rights reserved.
Keywords: Cathodic reactions; Cupric chloride; Sodium chloride; Chalcopyrite; Leaching of chalcopyrite; HydroCopper

1. Introduction
As sulfur dioxide emissions from the pyrometallurgical
industry are highly undesirable, the development of alternative hydrometallurgical process options for sulfide minerals is of great environmental importance. Chalcopyrite,
CuFeS2, is the most common copper mineral, available in
large quantities and with widespread distribution across
the globe. There is an incentive to develop economically
beneficial processes, like HydroCopper (Hyvrinen
et al., 2002), which convert the sulfur content of sulfide
minerals into elemental sulfur instead of sulfur dioxide.

This paper was presented at the EuroMaterials 2005 conference.


Corresponding author. Fax: +358 9 4512799.
E-mail address: Mari.Lundstrom@tkk.fi (M. Lundstrm).

0304-386X/$ - see front matter 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.hydromet.2006.07.002

Generally, the leaching of chalcopyrite in chloride


media, with Cu2+ as an oxidant, is more effective than
in sulfate media with Fe3+ as the oxidant. The redox
potential of the Cu2+/Cu+ couple is higher in chloride
solution compared to sulfate solution. Also the redox
potential of Cu2+/Cu+ is higher than that of ferric/
ferrous ions in concentrated chloride solutions. This
increase in the Cu2+/Cu+ redox potential is due to the
very strong complexation of cuprous ions with chloride
ions. Beside the rate of electron transfer, enhanced
crystallinity of the sulfur and lack of passivation support cupric chloride leaching of chalcopyrite. (Muir,
2002).
In chloride solution in the presence of cupric ions,
chalcopyrite is reported to dissolve according to reaction
(1) forming elemental sulfur (Habashi, 1978; Hietala

10

M. Lundstrm et al. / Hydrometallurgy 85 (2007) 916

and Hyvrinen, 2003; Hyvrinen et al., 2002; Olper and


Maccagni, 2003).

2
0
CuFeS2s 3Cu2
aq 4Cuaq Feaq 2Ss

Although cuprous chloride is generally not soluble in


water, in highly concentrated cupric chloride solutions,
high cuprous solubility can be achieved. When the cupric
chloride solution has a sodium chloride concentration of
250 g/l, reaction (1) can proceed already at the boiling point
of the solution at atmospheric pressure (Habashi, 1978).
In concentrated chloride solutions, copper ions readily
form complexes. Thus the leaching reaction can also be
suggested to progress in the presence of cupric and cuprous
complexes by reaction (2) (Wilson and Fisher, 1981).

CuFeS2s 3CuCl
aq 11Claq
0
4CuCl3 2
aq FeCl2aq 2Ss

Although a lot of hydrometallurgical studies in the


field of chalcopyrite leaching have been carried out, the
nature of leaching and rate limiting factors are not yet
totally understood. (Hackl et al., 1995; Munoz et al.,
1979; Parker et al., 2003; Roman and Benner, 1973).
In a previous study (Lundstrm et al., 2005) it was
shown that the process parameters can significantly affect
the dissolution rate of chalcopyrite and the properties of
the reaction product layers on the mineral. Increasing the
temperature from 70 to 90 C doubled the dissolution
current density of chalcopyrite. A critical cupric ion concentration of ca. 9 g/l was found, below which changes in
the cupric ion concentration did not significantly affect the
dissolution rate. The open circuit potential (OCP) increased with the increasing copper concentration and
followed the Nernst equation, with a slope of 60 mV/
decade. When the cupric concentration was greater than
9 g/l the increase in the cupric ion concentration was
observed to increase the dissolution rate of chalcopyrite.
The solution pH was shown to have a remarkable effect on
the composition of the reaction product layers as well as
on the dissolution rate. Quantitative calculations from the
SEM analysis and visual observations suggested that at
pH b 2.5 the reaction product layer on the chalcopyrite
surface was sulfur-rich and chalcopyrite was rapidly passivated. At higher pHs the reaction product layer was a
two-phase layer consisting mainly of iron and oxygen (in
the ratio 1:2 to 1:3). This was suggested to be a hydrated
iron oxide or goethite type layer.
It is clear that in the HydroCopper process the anodic
reaction is the dissolution of chalcopyrite. The dissolution
is either (i) under mixed or chemical control (with cupric
ion concentrations b9 g/l, pH = 2, T = 85 C) or (ii) under

diffusion control through a passivating product layer (with


cupric ion concentrations 9 g/l, pH= 2, T = 85 C,
diffusion coefficient, D, being 1.4 10 9 cm2/s) (Lundstrm et al., 2005). An experimental plot of the OCP vs.
cupric ion concentration was published earlier (Lundstrm
et al., 2005). Results showed that the OCP of chalcopyrite
in a concentrated chloride solution is in the range 450 to
600 mV vs. Ag/AgCl, depending on the process
parameters (temperature, cupric ion concentration and
pH). For example, with cupric ion concentrations 1 and
10 g/l, the OCP of CuFeS2 is ca. 510 and 570 mV vs. Ag/
AgCl, respectively (T = 90 C, pH= 2). It should be noted,
however, that the OCP of chalcopyrite in a concentrated
cupric chloride solution varies due to the non-homogeneity of the individual mineral samples.
To obtain a deeper knowledge about the leaching of
chalcopyrite, the present work focusses on the cathodic
reactions of the leach solution.
2. Materials
In the present work the concentration of NaCl used
was 280 g/l (4.8 M) and the cupric ion concentrations
were varied between 1 and 40 g/l. The temperature was
90 C and the pH was adjusted to 2 by HCl.
As chalcopyrite is a difficult material to make into
electrodes, Pt and glassy carbon (GC) were also used to
gain a clearer understanding of the cathodic reactions of
the cupric chloride solution. The chalcopyrite used was
from Pyhsalmi, Finland, its elemental composition
was analysed with SEM/EDS, the average wt.% was
Cu 30.0%, Fe 32.5%, S 36.3%, Si 0.5%, Al 0.4% and
Mg 0.3%. These values were near to the theoretical
composition of chalcopyrite (Cu 34.6%, Fe 30.4% and
S 34.9%). The chalcopyrite electrode surface was polished between every measurement with wetted grade 800
waterproof abrasive paper on a polishing wheel. After
polishing the chalcopyrite electrodes were rinsed with deionised water, then ethanol and then dried with a warm air
gun. The glassy carbon electrode, prepared at Outokumpu
Research (Pori, Finland), was a 5-mm disk glassy carbon
disk (Johnson Matthey Type 1, Alfa Aesar, Germany),
housed in a PTFE sheath.
A standard three-electrode electrochemical cell with a
thermostated water jacket was employed for the electrochemical measurements. In all measurements (except
rotating disk electrode, RDE) the cell was stirred with a
magnetic stirrer at 500 rpm. No purging of gases was
done. The counter electrode was a platinum sheet or wire,
the reference electrode was Ag/AgCl (REF201, Radiometer Analytical, France) placed in a sintered glass tube
containing a gel of agar powder, potassium chloride and

M. Lundstrm et al. / Hydrometallurgy 85 (2007) 916

distilled water. The Ag/AgCl reference electrode has a


potential of 197 mV vs. the standard hydrogen electrode
(SHE) (Bard and Faulkner, 2001). The reference electrode
junction was positioned in an external beaker and
connected to the cell via a salt bridge and 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) a 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.
3. Procedures
The RDE technique was used to study the mass transport as the Levich equation, Eq. (3), predicts the variation
in the transport-limited current as a function of the
electrode rotation rate
jlim 0:62zFD 2=3 v1=6 x1=2 C

where jlim is the limiting current density (A/cm2), z is the


number of electrons involved in the reaction, F is
Faraday's constant (96485 C/mol), D is the diffusion
coefficient (cm2/s), v is the kinematic viscosity (cm2/s),
is the rotation rate (1/s = 2f ) and C is the concentration
(mol/cm3). From the Levich equation the limiting current,
when plotted as a function of the square root of rotation
rate, 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.
On the basis of Eq. (3) it can be concluded (i) that jlim is
proportional to 1/2 when the concentration is constant
(Levich plot). On the other hand it follows also (ii) that at
constant , jlim is proportional to the concentration C.
Knowledge of these relationships can be applied to study
the reaction controlling mechanisms. The corresponding
unit rate constant, kT (cm/s), characterising the rate of
transport of a solution species to the RDE surface is given
by Eq. (4) (Gregory and Riddiford, 1956).
kT 0:62D 2=3 v1=6 x1=2

According to Gregory and Riddiford (1956) a discrepancy in the Levich equation, Eq. (3), lies in the constant
0.62, based on the I() value of 0.8934. A better estimation
of the diffusion coefficient is suggested to be given by Eq.
(5) for values of (D/v) in the range 04 10 3 cm2/s
(Gregory and Riddiford, 1956).
Il 0:8934 0:316D=v0:36

11

As the value of D 2/3 / I() remains constant, the correlation factor for the Levich equation can be taken into
account. Eq. (5) is known as the GregoryRiddiford
equation (MacHardy and Janssen, 2004).
Non-linearity of a Levich plot at low rotation rates may
be due to kinetic limitations. The KouteckyLevich equation, Eq. (6), (Bard and Faulkner, 2001) is applicable to
first order reactions and can be used to study the combined
effect of diffusion current and kinetic current (Lyons,
2002).
1=I 1=Ik 1=Id

Where I is the total current, Ik is the charge transfer


current and Id is the diffusion current. If Eq. (6) is valid,
the plot of I 1 vs. 1 gives a straight line and Ik can be
calculated from the y-axis intercept. An intercept near
the value of zero indicates a neglible effect of the charge
transfer process on the dissolution rate (Ohno et al.,
1990).
Cyclic voltammetry (CV) is a widely used electrochemical method and although it is not ideal for quantitative evaluation of system properties, it is however, a
powerful method, due to the ease of interpretation of
qualitative and semi-quantitative behaviour (Bard and
Faulkner, 2001). It can also provide rapid and reasonably
accurate determinations for reaction parameters, such as
rate constants (Nicholson, 1965). The peak current can be
represented as a function of diffusion coefficient, concentration and scan rate, Eq. (7), (Bard and Faulkner,
2001; Sundholm, 1987).
Ip 0:4463

zF3=2 A
RT 1=2

1=2

D0 C0 v

1=2

Where Ip is the peak height (A), A is the electrode


area (cm2), D0 is the diffusion coefficient of the reactive
species (cm2/s), C0 is the bulk concentration (mol/cm 3),
is the sweep rate (V/s), R is the gas constant (J/
mol K) and T is the temperature (K). According to Eq.
(7), if Ip is linearly dependent on the square root of the
sweep rate, the diffusion coefficient can be calculated.
This Eq. (7) at T = 25 C is also called the Randles
Sevcik equation, applicable for a reversible system, for
which the heterogeneous electron transfer reaction is
fast.
The difference between anodic peak potential (Epa) and
cathodic peak potential (Epc) is represented by Ep and
for a fully reversible reaction has a value of 59 / z mV
(T = 25 C) or 72 / z mV (T = 90 C) (Sundholm, 1987).
When the electron transfer reaction is quasi-reversible, the
rate of electron transfer is too slow to keep the redox
couple in equilibrium as the potential is changed. Thus

12

M. Lundstrm et al. / Hydrometallurgy 85 (2007) 916

evolution were observed. As the current densities measured


in sodium chloride solution (Figs. 1 and 2, lowest curve)
were low compared to those measured in cupric chloride
solutions (Fig. 2, three upper curves) similar to those used
in the HydroCopper process, it can be concluded that
neither platinum nor glassy carbon have any remarkable
reactions with pure sodium chloride electrolytes. Thus both
platinum and glassy carbon are suitable electrode materials
for the study of cathodic reactions in cupric chloride
solutions.
4.1. Cathodic reactions
Fig. 1. Third sweep from cyclic voltammograms with stationary platinum
and glassy carbon electrodes. [NaCl] = 280 g/l, sweep rate =50 mV/s,
pH= 2 and T = 90 C.

Ep increases with the scan rate () and can be used to


estimate the rate constant for electron transfer (k0) method
of Nicholson (Bard and Faulkner, 2001; Nicholson,
1965).
The Tafel method can be used to determine the value
of i0 from which the value of k0 can be calculated, Eq.
(8) (Bard and Faulkner, 2001).
i0 FAk 0 C

4. Results
The nature of cathodic reactions in the leaching process
was studied using platinum and glassy carbon electrodes.
Platinum is a strong adsorber and an excellent electrocatalyst for hydrogen evolution. The conventional voltammetric response of a polycrystalline platinum/aqueous
solution interface is usually divided into three sections: (i)
adsorption and desorption of hydrogen, b200 mV vs. Ag/
AgCl, (ii) double layer charging, 200 to 600 mV vs. Ag/
AgCl and (iii) oxygen adsorption and oxide formation,
600 to 1300 mV vs. Ag/AgCl. The shape, number and size
of the peaks caused by adsorption/desorption processes
depend on the crystal faces, electrode pre-treatment,
impurities present and the electrolyte (Bard and Faulkner,
2001; Casey, 1999; Sundholm, 1987).
Cyclic voltammetry (Fig. 1) was carried out with platinum and glassy carbon electrodes in a concentrated
sodium chloride solution ([NaCl] = 280 g/l) without cupric
ions. Neither platinum nor glassy carbon appeared to be
reactive in a concentrated sodium chloride solution (Fig. 1).
The electrochemical response of platinum suggested characteristic adsorption and desorption processes. At lower
and higher potentials high currents corresponding to gas

Cathodic polarization curves were measured with a


platinum RDE. The measurements were carried out at
90 C and pH 2, with 280 g/l of NaCl and the cupric ion
concentrations were 1, 10, 20, 30 or 40 g/l. The characteristic curve from the cathodic polarization measurements is similar to that of platinum shown in Fig. 2. The
curve begins from the OCP, reaching a limiting current (i),
which increases both with increasing cupric ion concentration and increasing RDE rotation rate. When the cupric
ion concentration is varied from 1 to 40 g/l, the redox
potential measured for the couple [CuCl]+ / [CuCl3]2 lies
in the range 600 to 700 mV vs. Ag/AgCl (Bonsdorff et al.,
2005). The thermodynamic calculations carried out with
the HSC Chemistry package, show much lower redox
potentials for [CuCl]+ / [CuCl3]2 (from 300 to 400 mV
vs. Ag/AgCl) (Lundstrm et al., 2005). This difference
can be explained by the limitations of the calculation
program database. On the grounds of the experiments
done and the thermodynamic considerations with the
HSC Chemistry package, it is suggested that when the

Fig. 2. Cathodic polarization curves for glassy carbon, platinum and


chalcopyrite RDEs (100 rpm). [NaCl] = 280 g/l, [Cu2+] = 20 g/l (for
three upper curves) and [Cu2+] = 0 g/l (for the lowest curve), sweep
rate = 0.7 mV/s, pH = 2 and T = 90 C.

M. Lundstrm et al. / Hydrometallurgy 85 (2007) 916

electrode is polarised from the OCP in the cathodic


direction, the complex [CuCl]+ is reduced to [CuCl3]2.
The reduction reaches a limiting current density at
approximately E = 300 mV vs. Ag/AgCl (i).
At ca. E = 250 mV vs. Ag/AgCl (Fig. 2) a small peak
(ii) is observed with lower, but not with higher rotation
rates, this peak is dependent on the initial cupric ion
concentration. Below E = 250 mV vs. Ag/AgCl the
current increases strongly (iii). Independent of whether or
not the peak was present, precipitated copper was always
observed on the platinum electrode surface after the
measurement. According to the Pourbaix diagram drawn
with the HSC Chemistry package, ([Cu2+] = 26.6 g/l,
[NaCl] = 250 g/l, T = 90 C) the precipitation of copper
begins at approximately E = 200 mV vs. Ag/AgCl. Thus
we propose that the peak (ii) is the reduction of the
complex [CuCl3]2 to metallic copper at ca. 250 mV vs.
Ag/AgCl. The reduction potential increases with the
increasing cupric ion concentration. Copper deposition is
followed by hydrogen evolution (iii) at potentials more
negative than 250 mV vs. Ag/AgCl.
4.2. The effect of mass transfer
In a concentrated cupric chloride electrolyte solution,
the shape of the polarization curve in the potential range
300 to 200 mV vs. Ag/AgCl (Fig. 2) shows a limiting
current density, which suggests that the cathodic
reaction ([CuCl]+ reduction to [CuCl3]2) is controlled
by mass transfer in this potential range.
The limiting current densities at E = 0 mV vs. Ag/
AgCl were taken from the cathodic polarisation curves
and plotted as Levich plots (Fig. 3) for cupric ion concentrations between 10 and 40 g/l. The response was
linear indicating a diffusion controlled process in the

Fig. 3. Levich plot with cupric ion concentrations 10, 20, 30 and 40 g/l.
[NaCl] = 280 g/l, pH = 2, T = 90 C. Limiting current density taken at
0 mV vs. Ag/AgCl.

13

Table 1
Diffusion coefficient of Cu2+ ion or complex calculated from the
experimental data
Method

[Cu2+]
(g/l)

Rotation
rate (rpm)

Levich D
(cm2/s)

GregoryRiddiford
D (cm2/s)

Levich plot
Levich plot
Levich plot
Levich plot
jlim[Cu2+]
jlim[Cu2+]
jlim[Cu2+]

40
30
20
10
140
140
140

1002500
1002500
1002500
1002500
100
900
2500

8.5 10 6
8.5 10 6
7.8 10 6
8.6 10 6
8.1 10 6
8.7 10 6
8.2 10 6

9.0 10 6
9.0 10 6
8.2 10 6
9.1 10 6
8.5 10 6
9.2 10 6
8.6 10 6

E = 0 mV vs. Ag/AgCl. [NaCl] = 280 g/l, pH = 2 and T = 90 C.

potential range investigated. Plots of [Cu2+] vs. jlim,


when E = 0 mV vs. Ag/AgCl, gave a linear response for
all the rotation rates studied: 100, 900 and 2500 rpm
(10.5, 94.2 and 261.8 Hz). Diffusion coefficients
calculated from the Levich plots and curves of [Cu2+]
vs. jlim are presented in Table 1, the mean value obtained
was 8.3 0.3 10 6 cm2/s.
The diffusion coefficients for Cu2+ calculated with the
GregoryRiddiford equation, Eq. (5), are also included in
Table 1. The mean diffusion coefficient was 8.8
0.4 10 6 cm2/s, i.e. approximately 5.5% more than
that calculated with the Levich method (Gregory and
Riddiford, 1956). The D values used in Gregory
Riddiford calculations were those shown in Table 1,
column 4. The kinematic viscosity was estimated as
4.55 10 3 cm2/s (Lobo and Quaresma, 1989).
The unit rate constant characterises the rate of transport
of the solution species to the rotating disk. In this case the
species are assumed to be cupric ions or complexes. The
values of the unit rate constants (kT), presented in Table 2,
show that kT is independent of the cupric ion concentration in the studied range and that kT varies as a function of
the rotation rate. The average unit rate constants are
suggested to be in the range 2.010.1 10 3 cm/s with
rotation rate from 100 to 2500 rpm.
When the diffusion coefficient of the cathodic reaction
(8.8 0.4 10 6 cm2/s) is compared to the value of
1.4 10 9 cm2/s measured earlier for the anodic reaction
i.e. chalcopyrite dissolution (Lundstrm et al., 2005), it is
clear that the cathodic values are significantly higher. This
suggests that mass transfer in the solution does not control
chalcopyrite dissolution in concentrated cupric chloride
solutions. However, the real HydroCopper process
does not operate at such low potentials (200 to 300 mV
vs. Ag/AgCl) in which the mass transfer is observed to
control the cathodic reaction (Fig. 2). Thus it is necessary
to study the reactions around the open circuit potential of
the concentrated cupric chloride solution.

14

M. Lundstrm et al. / Hydrometallurgy 85 (2007) 916

Table 2
Unit rate constants (kT, cm/s) of cupric ions/complexes calculated from D values (Levich plot) for different cupric concentrations with varying
rotation rate
Rotation rate
(rpm)
100
400
900
1600
2500

[Cu2+] = 10 g/l

[Cu2+] = 20 g/l

[Cu2+] = 30 g/l

[Cu2+] = 40 g/l

Average

kT (cm/s)

kT (cm/s)

kT (cm/s)

kT (cm/s)

kT (cm/s)

2.1 10
4.1 10 3
6.2 10 3
8.3 10 3
10.3 10 3

1.9 10
3.9 10 3
5.8 10 3
7.8 10 3
9.7 10 3

2.1 10
4.1 10 3
6.2 10 3
8.2 10 3
10.3 10 3

2.0 10
4.1 10 3
6.1 10 3
8.2 10 3
10.2 10 3

2.0 10 3
4.1 10 3
6.1 10 3
8.1 10 3
10.1 10 3

[NaCl] = 280 g/l, pH = 2 and T = 90 C.

4.3. The effect of charge transfer


Figs. 2 and 4 show that in the potential range near
the open circuit potential of the studied solution, the
cathodic reactions are not totally diffusion controlled.
This potential range is of the greater interest, when
considering the operation of the HydroCopper process. Thus it is important to understand the effect of
charge transfer on the cathodic reaction.
Cyclic voltammograms were measured with a glassy
carbon electrode to study the cathodic reaction in cupric
chloride solutions near the OCP of chalcopyrite in a
similar cupric chloride solution. Each measurement consisted of ten cycles in the potential range E = 200 to
900 mV vs. Ag/AgCl. The system achieved steady state
after six sweeps. These are presented in Fig. 4.
The peak currents of the anodic and cathodic reaction
in the solution were of the same magnitude and similar in
shape and area for the scan rates studied (10200 mV/s).
The average ratio of the anodic and cathodic peak currents
was 1.0 0.1 and was independent of the scan rate. Since
the anodic and cathodic peak currents were equal, the
oxidized and reduced forms of the ion or complex should

Fig. 4. Cyclic voltammetry as a function of scan rate (10, 20, 50, 100
and 200 mV/s) for stationary glassy carbon electrode (6th sweeps,
Eocp 900 mV 200 mV). Solution: [Cu2+] = 20 g/l, [NaCl] = 280 g/l,
pH = 2 and T = 90 C.

have similar diffusion coefficients. Both the anodic and


cathodic peak currents varied linearly with the square root
of the scan rate in the range 5 and 200 mV/s (correlation
coefficient N0.99 in all cases). D0, calculated from the
slope of the RandlesSevcik plot, was approximately
1.7 10 5 cm2/s being twice as much as the value presented in Table 1. That can be explained in Fig. 4, which
shows that the reaction does not represent a totally reversible system. The RandlesSevcik equation, Eq. (7), does
not hold when the value of Ep is not constant, hence
there is some error in D0 values calculated using this
method.
The peak-to-peak potential separation, Ep, for glassy
carbon in a cupric chloride solution ([Cu2+] = 20 g/l and
[NaCl] = 280 g/l) increased from 100 to 200 mV when the
sweep rate was increased from 10 to 200 mV/s. This
behaviour is characteristic of a quasi-reversible system,
i.e. the reaction kinetics at potentials close to the open
circuit potential are determined not only by diffusion but
also by the rate of electron transfer. The standard rate
constant for electron transfer, k0, was calculated from Ep
by Nicholson analysis (Nicholson, 1965) but did not

Fig. 5. Experimental (Pt, RDE) and simulated polarisation curves with


[Cu2+] = 10 and 20 g/l, [NaCl] = 280 g/l, T = 90 C, pH= 2, = 1600 rpm,
i0 = 8.3 10 4 A, D = 8.33 10 6 cm2/s and = 0.5.

M. Lundstrm et al. / Hydrometallurgy 85 (2007) 916

appear to give reliable k0 and i0 values for the temperatures used.


The approximation of the exchange current density ( j0)
value was obtained by the Tafel method from the cathodic
polarization curves measured with a platinum working
electrode. The exchange current density was independent
of the cupric ion concentrations in the range [Cu2+] = 10
40 g/l and gave an average value of 4.6 mA/cm2, corresponding to an exchange current (i0) value of 0.83 mA and
k0 value of 1.5 10 4 cm/s. The magnitude of k0 describes
the kinetic facility of the redox couple. The system is quasireversible, when 0.31/2 k0 2 10 51/2 cm/s (Bard and
Faulkner, 2001), at sweep rate 0.7 mV/s the corresponding
values are 7.9 10 3 k0 5.3 10 7 cm/s. In our system
k0 = 1.5 10 4 cm/s, confirming that the system studied is
quasi-reversible. It is clear that the j0 value for the cathodic
reaction (4.6 mA/cm2) is larger than that of the anodic
reaction (b 1.2 mA/cm2) (Lundstrm et al., 2005).
5. Simulation of polarization curves
If the system has no mass-transfer effects and the
surface concentrations do not differ appreciably from the
bulk values, then the current measured can be expressed
by the ButlerVolmer equation (Bard and Faulkner, 2001).
It is mentioned to be a good approximation for i values,
which are less than 10% of the limiting current. Under the
conditions that i =ilim the current follows the Levich
equation (Eq. (3)). In a system under mixed control, the
KouteckyLevich equation, Eq. (6), can be applied.
Levich plots at E=0 mV vs. Ag/AgCl indicated that the
system studied is diffusion controlled at relatively high
overpotentials (). The diffusion coefficient in the area of
limiting current density shown in Fig. 2 (i), was calculated to
be 8.3 0.310 6 cm2/s. Additionally, cyclic voltammetry
(Fig. 4) showed that at potentials near to the OCP, the overall
reaction rate is determined also by the rate of electron
transfer. As the system is quasi-reversible at lower cathodic
overpotentials and diffusion controlled at higher cathodic
overpotentials, the polarization curve can be simulated by
using the KouteckyLevich equation (Eq. (6)). Based on the
experimental data ([Cu2+] = 20 g/l, [NaCl] = 280 g/l,
T=90 C) the OCP was assumed to be 600 mV vs. Ag/
AgCl in all the simulated curves. Ik was calculated from the
ButlerVolmer equation (the transfer coefficient, , was
assumed to be 0.5 and i0 estimated with the Tafel method)
and Id from the Levich equation. Fig. 5 shows that the
experimentally measured and simulated polarisation curves
are in good agreement. Some deviation near the OCP can be
caused by the estimated value, which has an effect on the
shape of the curve. Also the i0 value, which is estimated by
the Tafel method, affects the same area. Furthermore, some

15

scatter in the plots is caused by variations in the OCP,


which depends on the cupric ion concentration and does
not actually stay constant in the simulated range. At high
current densities the simulation agrees well with the
experimental data. However, further studies are required
near the redox potential of the solution.
6. Conclusions
The aim of this study was to demonstrate the reduction
of copper ions in concentrated chloride electrolyte, similar
to that of the HydroCopper process solution. Platinum
and glassy carbon were found to be suitable electrode
materials for the study of cathodic reactions in cupric
chloride electrolytes. The redox system involving cupric
and cuprous complexes was shown to be quasi-reversible
around the open circuit potential of chalcopyrite.
The cathodic reactions under the conditions studied are
suggested to be (i) the reduction of [CuCl]+ to the
complex [CuCl3]2, (ii) the reduction of [CuCl3]2 to
solid copper at E 250 mV vs. Ag/AgCl and (iii) hydrogen evolution. From the cathodic polarization curves,
the diffusion coefficient for [CuCl]+ calculated by the
Levich method was 8.3 0.3 10 6 cm2 /s and the
corrected value obtained with the GregoryRiddiford
method was 8.8 0.4 10 6 cm2/s. The unit rate constant,
kT, was concluded to be independent of the cupric ion
concentration, but dependent on the RDE rotation rate.
The average unit rate constants are in the range 2.0
10.1 10 3 cm/s for rotation rate from 100 to 2500 rpm.
The cathodic reaction was shown not to limit the chalcopyrite dissolution in a HydroCopper type process
solution.
The exchange current density was estimated to be
4.6 mA/cm2. Simulated cathodic polarization curves, based
on the calculated values, were in good agreement with the
experimentally obtained curves, however, even small
changes in the value of or i0 have a remarkable effect
on the shape of the curve near the open circuit potential.
References
Bard, A.J., Faulkner, L.R., 2001. Electrochemical Methods Fundamentals and Applications. John Wiley and Sons Inc., 239,240,
335,347 pp.
Bonsdorff, R.v., et al., 2005. Electrochemical sensors for the HydroCopper process solution. Hydrometallurgy 77, 155161.
Casey, D.P., 1999. Anomalous redox behaviour of platinum electrodes
in aqueous media. M.Sc Thesis, National University of Ireland,
Cork, 148 pp.
Gregory, D.P., Riddiford, A.C., 1956. Transport to the surface of a
rotating disc. Journal of the Chemical Society 37563764.
Habashi, F., 1978. Chalcopyrite: its Chemistry and Metallurgy.
McGraw-Hill, Chatham. 165 pp.

16

M. Lundstrm et al. / Hydrometallurgy 85 (2007) 916

Hackl, R.P., Dreisinger, D.P., Peters, E., King, J.A., 1995. Passivation
of chalcopyrite during oxidative leaching in sulfate media. Hydrometallurgy 39 (1), 2548.
Hietala, K., Hyvrinen, O., 2003. HydroCopper a new
technology for copper production. Alta 2003 Copper Conference,
Perth, Australia, pp. 110.
Hyvrinen, O., Hmlinen, M., Leimala, R., 2002. Outokumpu
HydroCopper process a novel concept in copper production. In:
Peek, E., van Weert, G. (Eds.), Chloride Metallurgy 2002, 32nd
Annual Hydrometallurgy Meeting. MetSoc, Montreal, Quebec,
Canada, pp. 609612.
Lobo, V.M.M., Quaresma, J.L., 1989. Handbook of Electrolyte Solutions
Part B. Elsevier, Coimbra, Portugal, pp. 377, 1603, 1615.
Lundstrm, M., Aromaa, J., Forsn, O., Hyvrinen, O., Barker, M.H.,
2005. Leaching of chalcopyrite in cupric chloride solution.
Hydrometallurgy 77, 8995.
Lyons, M.E.G., 2002. Mediated electron transfer at redox active monolayers. Part 3: bimolecular outer-sphere, first order KouteckyLevich
and adduct formation mechanisms. Sensors 2, 473506.
MacHardy, S.J., Janssen, L.J.J., 2004. The diffusion coefficient of Cu(II)
ions in sulfuric acid-aqueous and methanesulfonic acid-methanol
solutions. Journal of Applied Electrochemistry 34, 169174.
Muir, D.M., 2002. Basic principles of chloride hydrometallurgy. In: Peek,
E., van Weert, G. (Eds.), Chloride Metallurgy 2002, 32nd Annual
Hydrometallurgy Meeting. MetSoc, Montreal, Canada, pp. 759778.

Munoz, P.B., Miller, J.D., Wadsworth, M.E., 1979. Reaction mechanism


for the acid ferric sulfate leaching of chalcopyrite. Metallurgical
Transactions. B, Process Metallurgy 10B, 149158 (June).
Nicholson, R.S., 1965. Theory and application of cyclic voltammetry
for measurement of electrode reaction kinetics. Analytical Chemistry 37, 13511355.
Ohno, H., Nishihara, H., Aramaki, K., 1990. The protection of iron
against corrosion with polymer films prepared by cathodic polymerization of halogenated xylenes. Corrosion Science 30 (6/7),
603615.
Olper, M., Maccagni, M., 2003. The modified Ecuprex process: a
promising hydrometallurgy approach for chalcopyrite-bearing
copper concentrates. In: Riveros, P.A., Dixon, D.G., Dreisinger,
D.B., Menacho, J.H. (Eds.), Copper 2003Cobre 2003. Met.Soc.,
Santiago, Chile, pp. 319334.
Parker, A., Klauber, C., Kougianos, A., Watling, H.R., van Broswijk, W.,
2003. An X-ray photoelectron spectroscopy study of the mechanism
of oxidative dissolution of chalcopyrite. Hydrometallurgy 71 (12),
265276.
Roman, R.J., Benner, B.R., 1973. The dissolution of copper concentrates. Minerals Science and Engineering 5 (1), 324.
Sundholm, G., 1987. Shkkemia 502 (Electrochemistry). Otakustantamo, Hmeenlinna, Finland. 240 pp.
Wilson, J.P., Fisher, W.W., 1981. Cupric chloride leaching of chalcopyrite. Journal of Metals 33 (2), 5257.

Das könnte Ihnen auch gefallen