You are on page 1of 9

Hydrometallurgy 93 (2008) 8896

Contents lists available at ScienceDirect

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

Leaching of chalcopyrite with ferric ion. Part II: Effect of redox potential
E.M. Crdoba a, J.A. Muoz b, M.L. Blzquez b, F. Gonzlez b, A. Ballester b,
a
b

Escuela de Ingeniera Metalrgica y Ciencia de los Materiales, Facultad de Ingenieras Fsico-Qumicas, Universidad Industrial de Santander, Bucaramanga, Colombia
Departamento de Ciencia de Materiales e Ingeniera Metalrgica, Facultad de Ciencias Qumicas, Universidad Complutense de Madrid, 28040 Madrid, Spain

A R T I C L E

I N F O

Available online 2 May 2008


Keywords:
Chalcopyrite
Ferric leaching
Redox potential

A B S T R A C T
This paper reports the effect of redox potential (or Fe3+/Fe2+ ratio) on chalcopyrite leaching. The relationship
between redox potential and other variables (iron concentration and temperature) is also evaluated. Leaching
tests were performed in stirred Erlenmeyer asks with 0.5 g of pure chalcopyrite and 100 mL of a Fe3+/Fe2+
sulphate solution. The redox potential ranged between 300 and 600 mV Ag/AgCl for the solution at a pH 1.8,
180 rpm, with temperatures at 35 C or 68 C. The results show that although ferric ion is responsible for the
oxidation of chalcopyrite, ferrous ion has an important role in that it controls precipitation and nucleation of
jarosites, which ultimately causes passivation of this sulphide. Chalcopyrite dissolves through the formation
of an intermediary product (covellite, CuS) that is later oxidized by ferric ion, releasing Cu2+ ions.
2008 Elsevier B.V. All rights reserved.

1. Introduction
Chalcopyrite, the most abundant mineral in copper ore bodies, is
also the most recalcitrant to hydrometallurgical processes (Dutrizac,
1978), and for that reason copper is mainly extracted by pyrometallurgy. However, the depletion of ore deposits and declining mineral
grades has encouraged the development of hydrometallurgical
processes for the treatment of chalcopyrite.
The dissolution of chalcopyrite by ferric ion is normally depicted by
the following chemical reactions (Dutrizac and MacDonald, 1974):
CuFeS2 4Fe3 YCu2 5Fe2 2S-

(Burkin, 1969), copper polysulphide with a decit of iron with respect


to chalcopyrite (Ammou-Chokroum et al., 1977 and Hackl et al., 1995),
or elemental sulphur (reaction (1)) (Muoz et al., 1979; Majima et al.,
1985 and Dutrizac, 1989).
In addition, several authors have concluded that the chalcopyrite
dissolution rate depends on the redox potential of the solution, the
best results being achieved under moderately oxidizing conditions
(Kametani and Aoki, 1985; Hiroyoshi et al., 2001; Okamoto et al.,
2003). Hiroyoshi et al. have proposed a model reaction involving the
intermediate reduction of chalcopyrite by ferrous ion to Cu2S and later
oxidation by ferric ion to release cupric ions:
CuFeS2 3Cu2 3Fe2 Y2Cu2 S 4Fe3 :

CuFeS2 4Fe3 3O2 2H2 OYCu2 5Fe2 2H2 SO4 :

The release of cuprous ion does not occur in a single step. A


Pourbaix diagram (Garrels and Christ, 1965) for the CuFeS2H2O
system would show the stability elds of several sulphides such as
Cu5FeS4, CuS or Cu2S between the stability elds of chalcopyrite and of
its dissolution.
After almost a century of research into the mechanisms of
chalcopyrite leaching, no unanimous theory has been proposed to
account for its slow kinetics. Nevertheless, there is consensus as to the
formation of a passivating layer on the chalcopyrite surface that slows
oxidation. The main theories were discussed in part I. Most of them
point to the formation of a diffusion layer surrounding the
chalcopyrite during dissolution, consisting of: bimetallic sulphide
Corresponding author. Tel.: +34 91 3944339; fax: +34 91 3944357.
E-mail address: ambape@quim.ucm.es (A. Ballester).
0304-386X/$ see front matter 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.hydromet.2008.04.016

Nicol and Lzaro (2003) have proposed a different mechanism to


explain the chalcopyrite dissolution at low redox potentials.
Kametani and Aoki identied a critical potential around 0.45 V vs.
SCE, associated with the onset of pyrite oxidation. A relatively low
redox potential has also been reported to have a favorable effect
during bioleaching (Barr et al., 1992). These authors observed low
redox potentials (400 to 450 mV SCE) after addition of ferrous ion,
which sped up the bioleaching kinetics of a copper concentrate. Third
et al. (2000, 2002) concluded that high ferric ion concentrations or
high redox potentials inhibited the bioleaching of chalcopyrite, but
they did not determine the cause of passivation.
Temperature is another factor which directly affects chalcopyrite
leaching rate. The high values of activation energy found by different
authors: 71 kJ/mol (Dutrizac et al., 1969), 84 kJ/mol (Muoz et al.,
1979), 88 kJ/mol (Hirato et al., 1987); clearly demonstrate the need of
high temperatures to break down bonds in the chalcopyrite crystal
lattice. In this study, the relationship between temperature and

E.M. Crdoba et al. / Hydrometallurgy 93 (2008) 8896

89

leaching kinetics is evaluated at 35 and 68 C, given that these


temperatures are appropriate in bioleaching processes with mesophilic and thermophilic bacteria respectively.
This research was prompted by the debate as to the relative
inuence of ferrous ion and of a low redox potential on chalcopyrite
dissolution. The specic aim was to study the role play by ferric and
ferrous ions, which control the potential, during leaching.
2. Materials and methods
2.1. Solids
All tests were performed with a chalcopyrite mineral (approximately
80% CuFeS2) from Messina, Transvaal (South Africa). Table 1 shows the
chemical composition of this mineral. The main impurities were pyrite
(FeS2), siderite (FeCO3) and silica (SiO2), as determined by X-ray analysis
(Fig. 1).
The mineral was dry-ground using a ball mill. The particle size
distribution was determined by laser pulse. The average particle size
was around 70 m. Additionally a BET surface area of 0.07 m2/g was
determined.
2.2. Leaching solutions
The different redox potentials of solutions were obtained by mixing
ferric and ferrous sulphates while keeping total iron concentration
constant at 5 g/L (in some tests 0.5 g/L) and pH at 1.8. Stock solutions of
ferric sulphate were prepared with 0 K (modied 9 K medium of
Silverman and Lundgren,1959) and Norris (Norris and Barr,1985) nutrient
mediums for the tests at 35 C and 68 C respectively, which are normally
used during bioleaching at these temperatures.

Fig. 1. X-ray diffractogram patter of the starting mineral.

At 35 C the effect of the initial redox potential was negligible, with


very low copper extractions (b2.5 %) in all cases (Fig. 2). The low
reactivity of the chalcopyrite surface is consistent with low consumption of the oxidizing agent (Fe3+).
The evolution of the redox potential with time shows that Fe3+/Fe2+
solutions tend towards equilibrium. This equilibrium potential, in a
range of 400500 mV vs. Ag/AgCl, is presumably related to the
standard potential of the Fe3+/Fe2+ couple. The value normally cited in
the literature is 771 mV vs. SHE or 564 mV vs. Ag/AgCl (Dry and
Bryson, 1988; Lowson, 1982); however, Cabral and Ignatiadis (2001)
obtained experimentally a lower value (644 mV vs. SHE or 437 mV vs.
Ag/AgCl) while assuming that the ratio of activity coefcients of both
ion species, Fe3+ and Fe2+, was equal to unity. This standard potential
value is within the equilibrium potential range observed in the
present study.

2.3. Leaching tests


All leaching tests were performed in an orbital shaker at 180 rpm and
35 or 68 C using 250 mL Erlenmeyer asks covered with hydrophobic
cotton so as to admit oxygen while reducing water loss through
evaporation. A low pulp density of 0.5% (100 mL of leaching solution
and 0.5 g of mineral) was chosen to avoid sharp changes in the redox
potential of the liquid medium during the onset of leaching.
Periodically, water evaporation was restored, pH adjusted when above
the initial value, redox potential recorded and 1 mL samples removed from
the liquid to obtain kinetic information on metal dissolution. Copper and
total iron concentration were determined by atomic absorption spectroscopy and ferrous ion concentration using a photocolorimetric method
based on the formation of a reddish colored complex of Fe2+ with orthophenantroline which was analyzed in an UVvis spectrophotometer at a
wavelength of 510 nm. Finally, solid residues were characterized by XRD
and SEM-EDS.
3. Results and discussion
3.1. Inuence of redox potential
The inuence of redox potential on the chalcopyrite dissolution rate
was tested at 35 C and 68 C. The redox potential values studied were 300,
400, 500 and 600 mV vs. Ag/AgCl.
Table 1
Chemical composition of the mineral tested
Element

Content (%)

Copper
Iron
Sulphur
Zinc
Lead

27.36
29.65
34.30
0.31
0.02

Fig. 2. Inuence of redox potential on the chalcopyrite leaching at 35 C and


[Fe]Total = 5 g/L.

90

E.M. Crdoba et al. / Hydrometallurgy 93 (2008) 8896

SEM examination of leaching residues did not show apparently


passivating layers on the chalcopyrite surface. The micrographs (Fig. 3)
show very clean surfaces with small amounts of a precipitate that is
insufcient to affect the reactivity of chalcopyrite. EDS microanalysis
of these residues are shown in Table 2. At low potential (400 mV), the
composition of the attacked chalcopyrite surface was practically
identical to the unattacked surface except for the presence of a small
oxidized precipitate. At high potential (600 mV), an oxidized jarositetype compound was detected on the mineral surface, presumably
potassium jarosite.
X-ray diffractograms of these residues (Fig. 4) show the presence of
several products: elemental sulphur (probably formed by reaction
(1)), covellite (formed possibly by chalcopyrite transformation, as will
be discussed later), goethite (only at low potential) and potassium
jarosite. The las two compounds are presumably formed by ferric ion
hydrolysis, according to the following reactions:
Fe2 SO4 3 4H2 OY2FeOOH 3H2 SO4

K 3Fe3 2SO2
4 6H2 OYKFe3 SO4 2 OH6 6H :

At high temperature (68 C), unlike at 35 C, the effect of the initial


redox potential was very pronounced (Fig. 5). At lower initial redox
potentials (Einitial 400 mV), the chalcopyrite dissolution rate was very
fast during the rst 5 days of leaching, and 80 or 90 % Cu extraction
was achieved (Fig. 5a). This successful leaching coincided with
potential values lower than approximately 450 mV (Fig. 5b) and
values of the Fe3+/Fe2+ ratio lower than unity (Fig. 5c). For times longer
than 5 days, the potential rose to 500 mV and the Fe3+/Fe2+ ratio
increased appreciably. Under these new conditions in the leaching
medium, chalcopyrite dissolution stopped.
These observations tend to support the hypothesis of Hiroyoshi
et al. (1997, 1999, 2000, 2001), according to which chalcopyrite
leaching with ferric sulphate is catalysed by ferrous ion. Furthermore,
in agreement with Hiroyoshi's results, there is a critical redox
potential value of 413 mV vs. Ag/AgCl above which chalcopyrite
dissolution slows down.

Fig. 3. SEM micrographs of the leaching residues at 35 C and [Fe]Total = 5 g/L:


a) Einitial = 400 mV and b) Einitial = 600 mV.

Table 2
EDS microanalysis of chalcopyrite leaching residues (results in weight percent)
Test

Chalcopyrite surface

Fe

Cu

35 C
400 mV
35 C
600 mV
68 C
400 mV
68 C
600 mV

Original
Attacked
Precipitate
Attacked
Precipitate
Attacked
Precipitate
Attacked
Precipitate

30.80
30.50
33.48
34.17
23.25
30.81
18.89
33.30
14.49

32.24
32.96
27.46
29.18
23.08
27.44
31.35
27.34
34.10

36.96
36.54
29.75
31.49
18.19
29.14
1.05
29.31
0.72

9.31
5.16
34.18
12.32
46.33
9.84
46.46

1.30
0.29
2.38
0.21
4.23

As noted earlier, various researchers have also observed better


chalcopyrite dissolution rates at low redox potentials at 30 C
(Okamoto et al., 2003) and at 90 C (Kametani and Aoki, 1985).
Kametani and Aoki determined a critical potential of 458 mV vs. Ag/
AgCl, and detected by XRD the presence of CuS in the leaching
residues, working at a very low potential (338 mV vs. Ag/AgCl).
SEM micrographs of the leaching residue at 68 C and an initial
redox potential of 400 mV (Fig. 6a) show chalcopyrite particles
surrounded by a precipitate identied by XRD as a mixture of
potassium jarosite, elemental sulphur and goethite (Fig. 7a). Although
this composition was similar to the leaching tests at 35 C, the amount
of precipitate was greater at 68 C. EDS microanalyses (Table 2) of that
leaching residue indicate an enrichment in potassium jarosite.
Under more oxidizing conditions (Einitial 500 mV), copper extractions were lower than 35 C (Fig. 5a). The low-magnication
backscattered electron micrograph of the residue at 600 mV
(Fig. 6b), shows few bright areas corresponding to clean chalcopyrite
surfaces and many dark areas corresponding to chalcopyrite covered

Fig. 4. X-ray diffractograms of the leaching residues at 35 C and [Fe]Total = 5 g/L:


a) Einitial = 400 mV and b) Einitial = 600 mV.

E.M. Crdoba et al. / Hydrometallurgy 93 (2008) 8896

91

Fig. 5. Inuence of redox potential on the chalcopyrite leaching at 68 C and [Fe]Total = 5 g/L.

by a diffusion layer. That lm formed on chalcopyrite particles is


clearly observed at higher magnication in Fig. 6c. the low porosity of
that layer would explain the poor results obtained at high redox

potentials. XRD analysis (Fig. 7b) and EDS microanalysis (Table 2)


point to that the diffusion layer consisted mainly of potassium jarosite.
Those differences were also related to a higher iron precipitation,
during the initial stages, at high than at low potentials (Fig. 5d).
Assuming that nucleation of jarosites on mineral particles (heterogeneous nucleation) was more important than nucleation out of
solution (homogeneous nucleation, that fast iron precipitation at high
potentials caused the rapid passivation of chalcopyrite.
At low potentials, the fast leaching kinetics initial stage was
followed by another very slow (Fig. 5a) and related to an abundant
iron precipitation (Fig. 5d). That would explain the presence of jarosite

Fig. 6. SEM micrographs of the leaching residues at 68 C and [Fe]Total =5 g/L: a)Einitial =400 mV,
b) Einitial =600 mV and c) circled particle in (b) at higher magnication.

Fig. 7. X-ray diffractograms of the leaching residues at 68 C and [Fe]Total = 5 g/L:


a) Einitial = 400 mV and b) Einitial = 600 mV.

92

E.M. Crdoba et al. / Hydrometallurgy 93 (2008) 8896

and goethite in the leaching residues at low potential (Fig. 6a), like in
experiments at high potentials (Fig. 6b). X-ray diffractograms of the
residues also show the presence of sulphur, although in a smaller
proportion than jarosite (Fig. 7).
The passivating nature of jarosites is in good agreement with
thermodynamic data. For the dissolution of jarosite at 25 C, Baron and
Palmer (1996) determined:
KFe3 SO4 2 OH6 6H YK 3Fe3 2SO2
4 6H2 O

11

the solubility product constant, Ksp = 10 , and the free energy of


jarosite formation, Gf, 3309.8 kJ/mol. These values are indicative of
the very low solubility and high stability of jarosites.
Casas et al. (2000) noted that jarosite formation in ferric sulphate
solutions starts from iron hydroxides, like Fe(OH)3, formed by
hydrolysis of ferric ion. The presence of goethite (FeOOH) in some
residues in the present study tends to conrm that hypothesis. The
action of goethite as activating agent in the formation of jarosite may
be depicted as follows:
2 Fe3 K 2SO2
4 FeOOHs
4H2 OYKFe3 SO4 2 OH6 3H :

These results point to that passivation of chalcopyrite during


leaching is due to the formation of a layer of jarosite which prevents
transportation of both electrons and ion species between the mineral
surface and the leaching medium.
Furthermore, the precipitation of iron, and hence the formation of
jarosites, is directly related to the redox potential of the solution. Thus,
redox potentials higher than the critical value (between 400 and
500 mV) favor Fe3+ precipitation as jarosite and the subsequent
chalcopyrite passivation.
The above observation is supported by Bigham et al. (1996) in that
the stability eld of potassium jarosite at pH lower than 2 is located at
potentials higher than 563 mV vs. Ag/AgCl. Thus, the closer the redox
potential of the leaching solution is to the beginning of the stability
eld of the jarosite, the faster it will precipitate.
As in tests at 35 C (Fig. 2), the kinetics curves at 68 C (Fig. 5) show
that the Fe3+/Fe2+ ratio or the redox potential of the leaching solution
tends asymptotically towards an equilibrium value of approximately
480 mV vs. Ag/AgCl. Only a few researchers have mentioned this trend
for leaching solutions of ferric sulphate. Of these, Dutrizac (1983)
noted that whereas the amount of jarosite formed increased with
ferric ion concentration, the Feprecipitated/Fedissolved ratio did not,
concluding that the solution tends toward equilibrium. Precipitation
of iron from ferric sulphate solutions, then, is difcult to control since
it occurs spontaneously until the system reaches equilibrium.
The leaching results at 68 C seem to demonstrate that precipitation and nucleation of jarosites on mineral particles would occur even
in tests with the fastest kinetics. Intermediate products were not
detected in the long-term tests, probably because of a jarosite lm
covering the particles. In order to generate residues to identify
possible intermediate products, new leaching tests were performed at
shorter times (1 h, 5 h and 1 day).
Figs. 8 and 9 show micrographs and X-ray diffractograms of
leaching residues at 68 C and Einitial = 400 mV with short attack-time.
Chalcopyrite surface transformations began at times as short as 1 h
(Fig. 8a). EDS microanalysis of precipitate shown in that gure
(named precipitate in Table 3) indicates an impoverishment of iron
and an enrichment of copper and sulphur. The XRD diffractogram of
the residue (Fig. 9a) evidences the formation of three products:
elemental sulphur, goethite and covellite. The last compound (CuS)
would be related to the EDS analysis of zone named precipitate in
Fig. 8a.
Micrographs of the 5 h leaching residue (Fig. 8b) show small
quantities of a precipitate over particles. EDS microanalyses (Table 3)

show again an enrichment of copper and sulphur on the chalcopyrite


surface and the presence of an oxidized compound in the zone named
precipitate, probably goethite. The diffractogram of that residue
(Fig. 9b) reveals the presence of the same three products found at 1 h:
elemental sulphur, goethite and covellite.
Finally, the growth of a product over the chalcopyrite surface
seems to start after 1 day of attack (Fig. 8c). EDS microanalyses
(Table 3) of that product indicate the presence of oxidized compounds
containing potassium and phosphorus in its composition. Two
additional products were detected by XRD (Fig. 9c): potassium jarosite
and ferric hydroxyphosphate.
The fact that goethite formed prior to the formation of jarosite
conrms the hypothesis that jarosite formation starts from Fe3+
hydro-oxidized compounds like goethite.
Even although the proportion of ferric hydroxyphosphate detected
in the residue after 1 day of leaching is small, like jarosite it can be
passivating. In fact, its chemical composition (Fe4(PO4)3(OH)3) may be
related to the alunite group (AB3(XO4)2(OH)6), as in the case of jarosite
(Lowson, 1982), where A stands for cations such as Na+, K+, H3O+, NH+4,
Pb2+ or Ag+; B stands for Fe3+ or Al3+, and XO4 usually for SO4, PO4 or
AsO4.
Another noteworthy nding was an increase of the CuS/chalcopyrite ratio, at least during the rst day of attack, as shown in the
diffractogram.

Fig. 8. SEM micrographs of the leaching residues at 68 C, [Fe]Total = 5 g/L, Einitial = 400 mV
and short times: a) 1 h, b) 5 h and c) 1 day.

E.M. Crdoba et al. / Hydrometallurgy 93 (2008) 8896

93

concluded that covellite forms either aerobically or anaerobically


through the reaction:
CuFeS2 Cu2 Y2CuS Fe2 :

12

Moreover, this transformation speeds up with temperature.


Kametani and Aoki, 1985, also found CuS in chalcopyrite leaching
residues carried out at a redox potential lower than 0.33 V vs. SCE.
All this suggests, then, that CuS only forms at low redox potentials.
Additionally, CuS formation can also occur during chalcopyrite
leaching by direct reaction between cupric ions and previouslyformed elemental sulphur, as follows:

3Cu2 4H2 O 4S-Y3CuS HSO


4 7H :

13

Nevertheless, the authors favor the idea that although this reaction
is thermodynamically possible, it does not occur because of the
hydrophobic nature of the elemental sulphur (Dutrizac and MacDonald, 1974).
Therefore, chemical reactions (8) and (12) become kinetically more
favorable.
3.2. Inuence of iron concentration

Fig. 9. X-ray diffractograms of the leaching residues at 68 C, [Fe]Total = 5 g/L,


Einitial = 400 mV and short times: a) 1 h, b) 5 h and c) 1 day.

The presence of CuS in the leaching products at short times


suggests that chalcopyrite dissolves in two steps. First it oxidizes,
forming CuS as an intermediary product:
CuFeS2 2Fe3 YCuS 3Fe2 S-:

Then, covellite is oxidized by ferric sulphate, releasing Cu2+ ions:


CuS 2Fe3 YCu2 2Fe2 S-

Also, we cannot rule out an initial reduction of chalcopyrite


(reaction (3)), as proposed by Hiroyoshi et al. (2001), followed by
oxidation of chalcocite to covellite by ferric sulphate through the
formation of copper-decient intermediate products, as reported by
Ferron (2003) in a study on the leaching of secondary copper minerals:
Cu2 S Fe2 SO4 3 YCuSO4 2FeSO4 CuS

10

The results reported in the previous section showed that


precipitation and nucleation of jarosites on mineral particles led to
passivation of chalcopyrite. In a rst attempt to explain and control
passivation, it was decided to study the effect of total iron
concentration.
Fig. 10 depicts the kinetics curves obtained during the leaching of
chalcopyrite at 68 C at two different iron concentrations: 0.5 and 5 g
FeTotal/L. Total iron concentration plays an important role in the
process: Reducing total iron concentration from 5 to 0.5 g Fe/L has a
negative effect on copper dissolution at low redox potential (400 mV)
and practically has no effect at high redox potential (600 mV).
The SEM study of the leaching residue at high potential and low
iron concentration showed chalcopyrite surfaces free of passivating
layers (Fig. 11). The X-ray diffractogram of this residue (Fig. 12)
conrmed the same compounds previously detected with 5 g/L of Fe
(Fig. 7b): S, jarosite, goethite and ferric hydroxyphosphate, the last
two being more important in this case.
These results indicate that in the Fe3+/ Fe2+ couple the ion closely
related to chalcopyrite dissolution is Fe3+. The role of Fe2+, then, would
be to achieve rapid equilibrium of the Fe3+/Fe2+ couple in solution,
thus controlling hydrolysis of the ferric ion, which could be ultimately
responsible for the passivation of chalcopyrite.
The positive effect observed in chalcopyrite leaching at 68 C when
iron concentration was increased from 0.5 to 5 g/L (or from 0.009 to
0.09 M) suggests that the process is at least partially controlled by the
diffusion of ions towards the chalcopyrite surface. That is consistent
with the ndings of Hirato et al. (1987) that the dissolution rate at
70 C increases with ferric ion concentration up to 0.1 M, while above
that value the enhancement is negligible.

Table 3
EDS microanalysis of chalcopyrite leaching residues at 68 C, [Fe]Total = 5 g/L,
Einitial = 400 mV and short times (results in weight percent)
Time
1h

Cu2 S Y Cu1;931;96 S Y Cu1;80 S Y CuS :

chalcocite

djujerite

digenite

covellite

11

Some researchers have detected the formation of covellite from


chalcopyrite under reducing conditions. Jang and Wadsworth, 1993,

5h
1 day

Chalcopyrite surface

Fe

Cu

Original
Attacked
Precipitate
Attacked
Precipitate
Attacked
Precipitate

30.80
30.16
34.06
37.32
28.99
32.84
12.39

32.24
31.99
8.87
29.64
28.76
29.74
23.91

36.96
36.33
48.76
33.04
30.38
33.05
4.45

3.92

0.65

1.52
8.31
11.87
4.37
54.68

94

E.M. Crdoba et al. / Hydrometallurgy 93 (2008) 8896

Fig. 12. X-ray diffractogram of the leaching residue at 68 C, Einitial = 600 mV and
[Fe]Total = 0.5 g/L.

Fig. 10. Inuence of iron concentration on the chalcopyrite leaching at 68 C.

3.3. Inuence of temperature


The effect of temperature was very pronounced, with negligible
copper dissolution at 35 C after 13 days of leaching (b3%) vs. almost
complete mineral dissolution at 68 C (Figs. 2 and 5). Thermal
activation therefore plays an important role in this process.
The effect of temperature in the leaching of chalcopyrite with ferric
sulphate was evaluated by calculating the activation energy in the
experimental temperature range of 35 C to 68 C. Four assays were
performed at 35, 46, 57 and 68 C, while all other experimental
conditions were constant, i.e.: 0.5% pulp density, Einitial = 400 mV and
[Fe]Total = 5 g/L.
The activation energy was calculated from the kinetic constants
and the semilogarithmic plot of the Arrhenius equation.
First, the fraction of reacted chalcopyrite was plotted vs. time for
each temperature (Fig. 13). Then, linear curves were plotted using the
simplied shrinking core model proposed by Sohn and Wadsworth
(1979), according to the following expression:
1  1  X 1=3 k  t

14

where X is the fraction of reacted chalcopyrite, k the kinetic constant


and t time. That equation is represented graphically in Fig. 14. The
slope of the nal experimental data for the kinetic curves at 57 C and

Fig. 11. SEM micrograph of the leaching residue at 68 C, Einitial = 600 mV and
[Fe]Total = 0.5 g/L.

68 C was zero (Fig. 13), probably due to the precipitation of jarosites,


and so they were removed from the linear regression.
According to Eq. (14), the slopes of the curves in Fig. 14 correspond
to the following kinetic constant values: 0.0006, 0.0044, 0.0319 and
0.0732 days 1 for 35, 46, 57 and 68 C respectively.
Most chemical reactions obey the Arrhenius equation (Logan,
2000; Levenspiel, 2004) and it was used to determine the effect of
temperature on the leaching of chalcopyrite:
k A  eEa =RT

15

where Ea is the activation energy, A the pre-exponential factor with


the same units as the kinetic constant k, R the universal gas constant
(8.314 J.K 1.mol 1) and T the absolute temperature (K). Then, the
logarithm of the kinetic constant is a linear function of the inverse of
temperature. Table 4 shows the values represented in the Arrhenius
plot in Fig. 15.
The activation energy for chalcopyrite dissolution, deduced from
the slope of the straight line of the Arrhenius plot, was appreciably
higher (130.7 kJ/mol) than that reported by other researchers (71
88 kJ/mol) in the range of temperature between 50 and 94 C and in
sulphate medium (Dutrizac et al. (1969), Muoz et al. (1979), Hirato
et al. (1987). Those differences could be attributed to the temperature
range assayed. The kinetic curves show that chalcopyrite dissolution is
negligible at 35 C, and that the energy barrier responsible for the slow
copper dissolution rate could be overcome at 68 C. However, the
enhancement of copper extraction is less pronounced above a certain
temperature (N50 C).
Based on the parameters established by Moore (1990) to elucidate
reaction mechanisms in different chemical processes, the high
activation energy value registered in this study (130.7 kJ/mol)
indicates that during the rst stage of chalcopyrite leaching, when
the kinetics is approximately linear, the system is under chemical
control. Therefore, the main drawback of chalcopyrite dissolution

Fig. 13. Inuence of temperature on the chalcopyrite leaching at Einitial = 400 mV.

E.M. Crdoba et al. / Hydrometallurgy 93 (2008) 8896

95

Fig. 15. Arrhenius plot.


Fig. 14. Variation of 1 (1 X)

1/3

over time.

would be the strong chemical bonding of the crystal lattice, which


requires a high input of energy to the system. These ndings support
the hypothesis of Hiskey (1993) that the transport of electrons
through vacants is very poor due to the n-type semiconductivity of
chalcopyrite, and hence the rst step during chalcopyrite oxidation is
the consumption of vacants to favor electron transportation through
the crystal lattice. Thus, a great deal of the energy applied to the
system in the form of heat is consumed by the displacement of ions
from the bulk of the particle to the surface, eliminating surface vacants
and favoring the transport of electrons through the chalcopyrite
surface.
At the same time, increasing iron concentration from 0.5 to 5 g/L
improves the leaching rate at 68 C (Fig. 10), and so the diffusion of
ferric ion to the chalcopyrite surface may also control the process.
Researchers have not reached a consensus on the control steps
governing chalcopyrite leaching. Our ndings are similar to those
reported by Linge (1976) and Hirato et al. (1987) with a linear branch
in the copper kinetic curve. Nevertheless, many researchers (Dutrizac
et al., 1969; Ferreira and Burkin, 1975; Muoz et al., 1979; Parker et al.,
1981; Kametani and Aoki, 1985; Hackl et al., 1995, etc.) have obtained
parabolic kinetics with an initial linear branch, similar to our results
but with very varied conclusions. This variety in the hypotheses to be
found in the literature reects a lack of agreement on the mechanisms
of chalcopyrite dissolution.
These results seem to indicate that hydrolysis and precipitation of
Fe3+ has a main role in chalcopyrite passivation in ferric solutions by
preventing the contact between the mineral surface ant the oxidizing
agent in solution.
Decreasing pH or removing some ions that favor iron precipitation
could be a way to prevent ferric ion hydrolysis. However, our attempts
in that direction (results not shown) have not solved the problem of
chalcopyrite passivation. Copper dissolution rate decreased when pH
was reduced from 0.5 to 2.0, perhaps because of that the species
responsible for the oxidation of chalcopyrite is not properly Fe3+ but
probably Fe(SO4)2 (Crdoba, 2005). Furthermore, the removal of
monovalent cations (K+, Na+ or NH+4) from solution neither prevents
hydrolysis and precipitation of iron as goethite and hydronium
jarosite that, like potassium jarosite, also tend to nucleate over
chalcopyrite particles.

308
319
330
341

k (days 1)
0.0006
0.0044
0.0319
00732

k (s 1)
9

6.9444 10
5.0926 10 8
3.6921 10 7
8.4722 10 7

FeCO3 H2 SO4 YFeSO4 CO2 H2 O:

16

In the case of pyrite, there are evidences that this sulphide only
dissolves when chalcopyrite is already passivated. Moreover, pyrite
rest potential is higher than that of chalcopyrite and, therefore, pyrite
enhances dissolution chalcopyrite through galvanic contact.
Finally, the comparison between studies on chalcopyrite leaching
is a difcult task because of differences in experimental conditions
used. However, unlike at high temperature, chalcopyrite dissolution
and ferric ion hydrolysis kinetics are very slow processes at low
temperature (35 C).
4. Conclusions
1. The redox potential is a key factor in the leaching of chalcopyrite. A
high potential at the onset of leaching provokes rapid passivation of
chalcopyrite.
2. Ferric/ferrous sulphate leaching solutions tend to reach equilibrium
when the activities of the two ions are equal, which is associated
with a critical potential of approximately 450 mV. When the redox
potential is very high initially, that tendency to equilibrium favors
rapid precipitation of ferric ion as jarosite and consequently
passivation of chalcopyrite.
3. The activation energy during chalcopyrite leaching was 130.7 kJ/
mol, which is a clear demonstration of the importance of thermal
activation in this process.
4. Increasing the iron concentration from 0.5 to 5 g/L had a positive
effect in the chalcopyrite leaching at 68 C.
5. Chalcopyrite dissolves through the intermediate formation of
covellite, CuS, which is later oxidized by ferric ion to release Cu2+
ions:
CuFeS2 2Fe3 YCuS 3Fe2 S-

CuS 2Fe3 YCu2 2Fe2 S-:

6. The elemental sulphur that forms during chalcopyrite leaching is


porous and does not form a passivating layer on the chalcopyrite
surface.

Table 4
Values represented in the Arrhenius plot of Fig. 15
T (K)

On the other hand, minority iron-bearing minerals in the starting


mineral (siderite and pyrite) could have an important role in leaching
kinetics. Siderite is quickly dissolved in acidic media, increasing the
iron concentration in solution:

1000 / T (K 1)

ln k (s 1)

3.25
3.13
3.03
2.93

18.79
16.79
14.81
13.98

References
Ammou-Chokroum, M., Cambazoglu, M., Steinmez, D., 1977. Oxydation menage de la
chalcopyrite en solution acide: analyses cintique de ractions. II. Modles
diffusionales. Bulletin de la Societe francaise de mineralogie et de
cristallographie 100, 161177.

96

E.M. Crdoba et al. / Hydrometallurgy 93 (2008) 8896

Baron, D., Palmer, C.D., 1996. Solubility of jarosite at 435 C. Geochimica et


Cosmochimica Acta 60 (2), 185195.
Barr, D.W., Jordan, M.A., Norris, P.R., Phillips, C.V., 1992. An investigation into bacterial
cell, ferrous iron, pH and Eh interactions during thermophilic leaching of copper
concentrates. Minerals Engineering 5 (35), 557567.
Bigham, J.M., Schwertmann, U., Traina, S.J., Winland, R.L., Wolf, M., 1996. Schwertmannite and the chemical modeling of iron in acid sulfate waters. Geochimica et
Cosmochimica Acta 60 (12), 21112121.
Burkin, A.R., 1969. Solid-state transformations during leaching. Minerals science and
engineering 1 (1), 414.
Cabral, T., Ignatiadis, I., 2001. Mechanistic study of the pyritesolution interface during
the oxidative bacterial dissolution of pyrite (FeS2) by using electrochemical
techniques. International journal of mineral processing 62, 4164.
Casas, J.M., Lienqueo, M.E., Cubillos, F., Herrera, L., 2000. Modelacin cintica de la
precipitacin de hierro como jarosita en soluciones lixiviantes utilizando la bacteria
Thiobacillus ferrooxidans. Congreso Chileno de Ingeniera Qumica. Universidad de
Santiago. Octubre de 2000.
Crdoba, E.M. (2005). Nuevas evidencias sobre los mecanismos de lixiviacin y
biolgica de la calcopirita. Tesis Doctoral. Departamento de Ciencia de los Materiales
e Ingeniera Metalrgica. Facultad de Ciencias Qumicas. Universidad Complutense
de Madrid.
Dry, M.J., Bryson, A.W., 1988. Prediction of redox potential in concentrated iron sulphate
solutions. Hydrometallurgy 21, 5972.
Dutrizac, J.E., 1978. The kinetics of dissolution of chalcopyrite in ferric ion media.
Metallurgical Transactions B 9B, 431439.
Dutrizac, J.E., 1983. Factors affecting alkali jarosite precipitation. Metallurgical
Transactions B 14B, 531539.
Dutrizac, J.E., 1989. Elemental sulphur formation during the ferric sulphate leaching of
chalcopyrite. Canadian Metallurgical Quarterly 28 (4), 337344.
Dutrizac, J.E., MacDonald, R.J.C., 1974. Ferric ion as a leaching medium. Minerals science
and engineering 6 (2), 59100.
Dutrizac, J.E., MacDonald, R.J.C., Ingraham, T.R., 1969. The kinetics of dissolution of
synthetic chalcopyrite in aqueous acidic ferric sulfate solutions. Transactions of the
Metallurgical Society of AIME 245, 955959.
Ferreira, R.C.H., Burkin, A.R., 1975. Acid leaching of chalcopyrite. In: Bulkies, A.R. (Ed.),
Leaching and Reduction in Hydrometallurgy. Inst. Min. Met. , pp. 5456. Londres.
Ferron, C.J., 2003. Leaching of secondary copper minerals using regenerated ferric
sulphate. In: Riveros, P.A., Dixon, D., Dreisinger, D.B., Menacho, J. (Eds.), COPPER
2003, Vol. VI Hydrometallurgy of Copper (Book 1), pp. 337352. Santiago, Chile.
Garrels, R.M., Christ, C.L., 1965. Solution, Minerals, and Equilibria. Editorial Harper &
Row, New York, pp. 213233.
Hackl, R.P., Dreisinger, D.B., Peters, E., King, J.A., 1995. Passivation of chalcopyrite during
oxidative leaching in sulfate media. Hydrometallurgy 39, 2548.
Hirato, T., Majima, H., Awakura, Y., 1987. The leaching of chalcopyrite with ferric sulfate.
Metallurgical Transactions B 18B, 489496.
Hiroyoshi, N., Hirota, M., Hirajima, T., Tsunekawa, M., 1997. A case of ferrous sulfate
addition enhancing chalcopyrite leaching. Hydrometallurgy 47, 3745.
Hiroyoshi, N., Hirota, M., Hirajama, T., Tsunekawa, M., 1999. Inhibitory effect of ironoxidizing bacteria on ferrous-promoted chalcopyrite leaching. Biotechnology and
Bioengineering 64 (4), 478483.
Hiroyoshi, N., Miki, H., Hirajima, T., Tsunekawa, M., 2000. A model for ferrous-promoted
chalcopyrite leaching. Hydrometallurgy 57, 3138.
Hiroyoshi, N., Miki, H., Hirajima, T., Tsunekawa, M., 2001. Enhancement of chalcopyrite
leaching by ferrous ions in acidic ferric sulfate solutions. Hydrometallurgy 60,
185197.

Hiskey, J.B., 1993. Chalcopyrite semiconductor electrochemistry and dissolution. In:


Reddy, R.G., Weizenbach, R.N. (Eds.), The Paul E. Queneau International Symposium,
Extractive Metallurgy of Copper, Nickel and Cobalt. Volume I: Fundamental Aspects.
The Minerals, Metals & Materials Society, pp. 949969.
Jang, J.H., Wadsworth, M.E., 1993. Hydrothermal conversion of chalcopyrite under
controlled Eh and pH. In: Reddy, R.G., Weizenbach, R.N. (Eds.), The Paul E.
Queneau International Symposium, Extractive Metallurgy of Copper, Nickel and
Cobalt. Volume I: Fundamental Aspects. The Minerals, Metals & Materials Society,
pp. 689707.
Kametani, H., Aoki, A., 1985. Effect of suspension potential on the oxidation rate of
copper concentrate in a sulfuric acid solution. Metallurgical Transactions B 16B,
695705.
Levenspiel, O., 2004. Captulo 3: Interpretacin de los datos obtenidos en un reactor
intermitente, Ingeniera de las reacciones qumicas, 3a edicin. Editorial Limusa
Wiley, Mxico, pp. 3867.
Linge, H.G., 1976. A study of chalcopyrite dissolution in acid ferric nitrate by
potentiometric titration. Hydrometallurgy 2, 5164.
Logan, S.R., 2000. Captulo 1: Fundamentos empricos de la cintica qumica. Fundamentos de cintica y qumica. Editorial Addison Wesley, Madrid, pp. 122.
Lowson, R., 1982. Aqueous oxidation of pyrite by molecular oxygen. Chemical Reviews
82 (5), 461497.
Majima, H., Awakura, Y., Hirato, T., Tanaka, T., 1985. The leaching of chalcopyrite in ferric
chloride and ferric sulfate solutions. Canadian Metallurgical Quarterly 24 (4),
283291.
Moore, J.J. (1990). Chapter 3: Reaction kinetics. Chemical Metallurgy. Ed. Butterworth
Heinemann. 2nd edition, Great Britain, 95126.
Muoz, P.B., Miller, J.D., Wadsworth, M.E., 1979. Reaction mechanism for the acid ferric
sulfate leaching of chalcopyrite. Metallurgical Transaction B 10B, 149158.
Nicol, M.J., Lzaro, I., 2003. The role of non-oxidative processes in the leaching of
chalcopyrite. In: Riveros, P.A., Dixon, D., Dreisinger, D.B., Menacho, J. (Eds.), COPPER
2003. Volume VIHydrometallurgy of Copper (Book 1), pp. 367381. Santiago,
Chile.
Norris, P.R., Barr, D.W., 1985. Growth and iron oxidation by acidophilic moderate
thermophiles. FEMS microbiology letters 28, 221224.
Okamoto, H., Nakayama, R., Tunekawa, M., Hiroyoshi, N., 2003. Improvement of
chalcopyrite leaching in acidic sulfate solutions by redox potential control. In:
Riveros, P.A., Dixon, D., Dresinger, D.B., Menacho, J. (Eds.), COPPER 2003. Volume VI
Hydrometallurgy of Copper (Book 1), pp. 6781. Santiago, Chile.
Parker, A.J., Paul, R.L., Power, G.P., 1981. Electrochemical aspects of leaching copper from
chalcopyrite in ferric and cupric salt solutions. Australian journal of chemistry 34,
1334.
Silverman, M.P., Lundgren, D.G., 1959. Studies on the chemoautotrophic iron bacterium
Ferrobacillus ferrooxidans. An improved medium and a harvesting procedure for
securing high cell yields. Journal of bacteriology 77, 642.
Sohn, H.Y., Wadsworth, M.E., 1979. Rate Processes of Extractive Metallurgy. Plenum,
New York, p. 133.
Third, K.A., Cord-Ruwisch, R., Watling, H.R., 2000. The role of iron-oxidizing bacteria in
stimulation or inhibition of chalcopyrite bioleaching. Hydrometallurgy 57,
225233.
Third, K.A., Cord-Ruwisch, R., Watling, H.R., 2002. Control of the redox potential by
oxygen limitation improves bacterial leaching of chalcopyrite. Biotechnology and
Bioengineering 78 (4), 433441.

You might also like