Anda di halaman 1dari 9

Hydrometallurgy 93 (2008) 88–96

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. Córdoba a, J.A. Muñoz b, M.L. Blázquez b, F. González b, A. Ballester b,⁎
a
Escuela de Ingeniería Metalúrgica y Ciencia de los Materiales, Facultad de Ingenierías Físico-Químicas, Universidad Industrial de Santander, Bucaramanga, Colombia
b
Departamento de Ciencia de Materiales e Ingeniería Metalúrgica, Facultad de Ciencias Químicas, Universidad Complutense de Madrid, 28040 Madrid, Spain

A R T I C L E I N F O A B S T R A C T

Available online 2 May 2008 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
Keywords: tests were performed in stirred Erlenmeyer flasks with 0.5 g of pure chalcopyrite and 100 mL of a Fe3+/Fe2+
Chalcopyrite sulphate solution. The redox potential ranged between 300 and 600 mV Ag/AgCl for the solution at a pH 1.8,
Ferric leaching
180 rpm, with temperatures at 35 °C or 68 °C. The results show that although ferric ion is responsible for the
Redox potential
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 (Burkin, 1969), copper polysulphide with a deficit of iron with respect
to chalcopyrite (Ammou-Chokroum et al., 1977 and Hackl et al., 1995),
Chalcopyrite, the most abundant mineral in copper ore bodies, is or elemental sulphur (reaction (1)) (Muñoz et al., 1979; Majima et al.,
also the most recalcitrant to hydrometallurgical processes (Dutrizac, 1985 and Dutrizac, 1989).
1978), and for that reason copper is mainly extracted by pyrome- In addition, several authors have concluded that the chalcopyrite
tallurgy. However, the depletion of ore deposits and declining mineral dissolution rate depends on the redox potential of the solution, the
grades has encouraged the development of hydrometallurgical best results being achieved under moderately oxidizing conditions
processes for the treatment of chalcopyrite. (Kametani and Aoki, 1985; Hiroyoshi et al., 2001; Okamoto et al.,
The dissolution of chalcopyrite by ferric ion is normally depicted by 2003). Hiroyoshi et al. have proposed a model reaction involving the
the following chemical reactions (Dutrizac and MacDonald, 1974): intermediate reduction of chalcopyrite by ferrous ion to Cu2S and later
oxidation by ferric ion to release cupric ions:
CuFeS2 þ 4Fe3þ YCu2þ þ5Fe2þ þ 2S- ð1Þ
CuFeS2 þ 3Cu2þ þ 3Fe2þ Y2Cu2 S þ 4Fe3þ : ð3Þ

CuFeS2 þ 4Fe3þ þ 3O2 þ 2H2 OYCu2þ þ 5Fe2þ þ 2H2 SO4 : ð2Þ Nicol and Lázaro (2003) have proposed a different mechanism to
explain the chalcopyrite dissolution at low redox potentials.
The release of cuprous ion does not occur in a single step. A Kametani and Aoki identified a critical potential around 0.45 V vs.
Pourbaix diagram (Garrels and Christ, 1965) for the CuFeS2–H2O SCE, associated with the onset of pyrite oxidation. A relatively low
system would show the stability fields of several sulphides such as redox potential has also been reported to have a favorable effect
Cu5FeS4, CuS or Cu2S between the stability fields of chalcopyrite and of during bioleaching (Barr et al., 1992). These authors observed low
its dissolution. redox potentials (400 to 450 mV SCE) after addition of ferrous ion,
After almost a century of research into the mechanisms of which sped up the bioleaching kinetics of a copper concentrate. Third
chalcopyrite leaching, no unanimous theory has been proposed to et al. (2000, 2002) concluded that high ferric ion concentrations or
account for its slow kinetics. Nevertheless, there is consensus as to the high redox potentials inhibited the bioleaching of chalcopyrite, but
formation of a passivating layer on the chalcopyrite surface that slows they did not determine the cause of passivation.
oxidation. The main theories were discussed in part I. Most of them Temperature is another factor which directly affects chalcopyrite
point to the formation of a diffusion layer surrounding the leaching rate. The high values of activation energy found by different
chalcopyrite during dissolution, consisting of: bimetallic sulphide authors: 71 kJ/mol (Dutrizac et al., 1969), 84 kJ/mol (Muñoz et al.,
1979), 88 kJ/mol (Hirato et al., 1987); clearly demonstrate the need of
⁎ Corresponding author. Tel.: +34 91 3944339; fax: +34 91 3944357. high temperatures to break down bonds in the chalcopyrite crystal
E-mail address: ambape@quim.ucm.es (A. Ballester). lattice. In this study, the relationship between temperature and

0304-386X/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.hydromet.2008.04.016
E.M. Córdoba et al. / Hydrometallurgy 93 (2008) 88–96 89

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


temperatures are appropriate in bioleaching processes with meso-
philic and thermophilic bacteria respectively.
This research was prompted by the debate as to the relative
influence of ferrous ion and of a low redox potential on chalcopyrite
dissolution. The specific 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 Fig. 1. X-ray diffractogram patter of the starting mineral.
(FeS2), siderite (FeCO3) and silica (SiO2), as determined by X-ray analysis
(Fig. 1). At 35 °C the effect of the initial redox potential was negligible, with
The mineral was dry-ground using a ball mill. The particle size very low copper extractions (b2.5 %) in all cases (Fig. 2). The low
distribution was determined by laser pulse. The average particle size reactivity of the chalcopyrite surface is consistent with low consump-
was around 70 μm. Additionally a BET surface area of 0.07 m2/g was tion of the oxidizing agent (Fe3+).
determined. The evolution of the redox potential with time shows that Fe3+/Fe2+
solutions tend towards equilibrium. This equilibrium potential, in a
2.2. Leaching solutions range of 400–500 mV vs. Ag/AgCl, is presumably related to the
standard potential of the Fe3+/Fe2+ couple. The value normally cited in
The different redox potentials of solutions were obtained by mixing the literature is 771 mV vs. SHE or 564 mV vs. Ag/AgCl (Dry and
ferric and ferrous sulphates while keeping total iron concentration Bryson, 1988; Lowson, 1982); however, Cabral and Ignatiadis (2001)
constant at 5 g/L (in some tests 0.5 g/L) and pH at 1.8. Stock solutions of obtained experimentally a lower value (644 mV vs. SHE or 437 mV vs.
ferric sulphate were prepared with 0 K (modified 9 K medium of Ag/AgCl) while assuming that the ratio of activity coefficients of both
Silverman and Lundgren,1959) and Norris (Norris and Barr,1985) nutrient ion species, Fe3+ and Fe2+, was equal to unity. This standard potential
mediums for the tests at 35 °C and 68 °C respectively, which are normally value is within the equilibrium potential range observed in the
used during bioleaching at these temperatures. 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 flasks 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 spectro-
scopy and ferrous ion concentration using a photocolorimetric method
based on the formation of a reddish colored complex of Fe2+ with ortho-
phenantroline which was analyzed in an UV–vis spectrophotometer at a
wavelength of 510 nm. Finally, solid residues were characterized by XRD
and SEM-EDS.

3. Results and discussion

3.1. Influence of redox potential

The influence 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 27.36
Iron 29.65
Sulphur 34.30
Zinc 0.31
Fig. 2. Influence of redox potential on the chalcopyrite leaching at 35 °C and
Lead 0.02
[Fe]Total = 5 g/L.
90 E.M. Córdoba et al. / Hydrometallurgy 93 (2008) 88–96

SEM examination of leaching residues did not show apparently Table 2


passivating layers on the chalcopyrite surface. The micrographs (Fig. 3) EDS microanalysis of chalcopyrite leaching residues (results in weight percent)

show very clean surfaces with small amounts of a precipitate that is Test Chalcopyrite surface S Fe Cu O K
insufficient to affect the reactivity of chalcopyrite. EDS microanalysis Original 30.80 32.24 36.96
of these residues are shown in Table 2. At low potential (400 mV), the 35 °C Attacked 30.50 32.96 36.54
composition of the attacked chalcopyrite surface was practically 400 mV Precipitate 33.48 27.46 29.75 9.31
35 °C Attacked 34.17 29.18 31.49 5.16
identical to the unattacked surface except for the presence of a small
600 mV Precipitate 23.25 23.08 18.19 34.18 1.30
oxidized precipitate. At high potential (600 mV), an oxidized jarosite- 68 °C Attacked 30.81 27.44 29.14 12.32 0.29
type compound was detected on the mineral surface, presumably 400 mV Precipitate 18.89 31.35 1.05 46.33 2.38
potassium jarosite. 68 °C Attacked 33.30 27.34 29.31 9.84 0.21
X-ray diffractograms of these residues (Fig. 4) show the presence of 600 mV Precipitate 14.49 34.10 0.72 46.46 4.23

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 As noted earlier, various researchers have also observed better
hydrolysis, according to the following reactions: chalcopyrite dissolution rates at low redox potentials at 30 °C
(Okamoto et al., 2003) and at 90 °C (Kametani and Aoki, 1985).
Fe2 ðSO4 Þ3 þ4H2 OY2FeOOH þ 3H2 SO4 ð4Þ 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).
Kþ þ 3Fe3þ þ 2SO2 þ
4 þ 6H2 OYKFe3 ðSO4 Þ2 ðOHÞ6 þ6H : ð5Þ SEM micrographs of the leaching residue at 68 °C and an initial
At high temperature (68 °C), unlike at 35 °C, the effect of the initial redox potential of 400 mV (Fig. 6a) show chalcopyrite particles
redox potential was very pronounced (Fig. 5). At lower initial redox surrounded by a precipitate identified by XRD as a mixture of
potentials (Einitial ≤ 400 mV), the chalcopyrite dissolution rate was very potassium jarosite, elemental sulphur and goethite (Fig. 7a). Although
fast during the first 5 days of leaching, and 80 or 90 % Cu extraction this composition was similar to the leaching tests at 35 °C, the amount
was achieved (Fig. 5a). This successful leaching coincided with of precipitate was greater at 68 °C. EDS microanalyses (Table 2) of that
potential values lower than approximately 450 mV (Fig. 5b) and leaching residue indicate an enrichment in potassium jarosite.
values of the Fe3+/Fe2+ ratio lower than unity (Fig. 5c). For times longer Under more oxidizing conditions (Einitial ≥ 500 mV), copper extrac-
than 5 days, the potential rose to 500 mV and the Fe3+/Fe2+ ratio tions were lower than 35 °C (Fig. 5a). The low-magnification
increased appreciably. Under these new conditions in the leaching backscattered electron micrograph of the residue at 600 mV
medium, chalcopyrite dissolution stopped. (Fig. 6b), shows few bright areas corresponding to clean chalcopyrite
These observations tend to support the hypothesis of Hiroyoshi surfaces and many dark areas corresponding to chalcopyrite covered
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: 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. a) Einitial = 400 mV and b) Einitial = 600 mV.
E.M. Córdoba et al. / Hydrometallurgy 93 (2008) 88–96 91

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

by a diffusion layer. That film formed on chalcopyrite particles is potentials. XRD analysis (Fig. 7b) and EDS microanalysis (Table 2)
clearly observed at higher magnification in Fig. 6c. the low porosity of point to that the diffusion layer consisted mainly of potassium jarosite.
that layer would explain the poor results obtained at high redox 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 (hetero-
geneous 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, Fig. 7. X-ray diffractograms of the leaching residues at 68 °C and [Fe]Total = 5 g/L:
b) Einitial =600 mV and c) circled particle in (b) at higher magnification. a) Einitial = 400 mV and b) Einitial = 600 mV.
92 E.M. Córdoba et al. / Hydrometallurgy 93 (2008) 88–96

and goethite in the leaching residues at low potential (Fig. 6a), like in show again an enrichment of copper and sulphur on the chalcopyrite
experiments at high potentials (Fig. 6b). X-ray diffractograms of the surface and the presence of an oxidized compound in the zone named
residues also show the presence of sulphur, although in a smaller “precipitate”, probably goethite. The diffractogram of that residue
proportion than jarosite (Fig. 7). (Fig. 9b) reveals the presence of the same three products found at 1 h:
The passivating nature of jarosites is in good agreement with elemental sulphur, goethite and covellite.
thermodynamic data. For the dissolution of jarosite at 25 °C, Baron and Finally, the growth of a product over the chalcopyrite surface
Palmer (1996) determined: seems to start after 1 day of attack (Fig. 8c). EDS microanalyses
(Table 3) of that product indicate the presence of oxidized compounds
KFe3 ðSO4 Þ2 ðOHÞ6 þ6Hþ YKþ þ 3Fe3þ þ 2SO2
4 þ 6H2 O ð6Þ containing potassium and phosphorus in its composition. Two
− 11 additional products were detected by XRD (Fig. 9c): potassium jarosite
the solubility product constant, Ksp = 10 , and the free energy of
and ferric hydroxyphosphate.
jarosite formation, ΔG°f, − 3309.8 kJ/mol. These values are indicative of
The fact that goethite formed prior to the formation of jarosite
the very low solubility and high stability of jarosites.
confirms the hypothesis that jarosite formation starts from Fe3+
Casas et al. (2000) noted that jarosite formation in ferric sulphate
hydro-oxidized compounds like goethite.
solutions starts from iron hydroxides, like Fe(OH)3, formed by
Even although the proportion of ferric hydroxyphosphate detected
hydrolysis of ferric ion. The presence of goethite (FeOOH) in some
in the residue after 1 day of leaching is small, like jarosite it can be
residues in the present study tends to confirm that hypothesis. The
passivating. In fact, its chemical composition (Fe4(PO4)3(OH)3) may be
action of goethite as activating agent in the formation of jarosite may
related to the alunite group (AB3(XO4)2(OH)6), as in the case of jarosite
be depicted as follows:
(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
2 Fe3þ þ Kþ þ 2SO2
4 þ FeOOHðsÞ
þ 4H2 OYKFe3 ðSO4 Þ2 ðOHÞ6 þ3Hþ : ð7Þ AsO4.
Another noteworthy finding was an increase of the CuS/chalcopyr-
ite ratio, at least during the first day of attack, as shown in the
These results point to that passivation of chalcopyrite during diffractogram.
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 field 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
field 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 difficult to control since
it occurs spontaneously until the system reaches equilibrium.
The leaching results at 68 °C seem to demonstrate that precipita-
tion 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 film
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 figure
(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 Fig. 8. SEM micrographs of the leaching residues at 68 °C, [Fe]Total = 5 g/L, Einitial = 400 mV
quantities of a precipitate over particles. EDS microanalyses (Table 3) and short times: a) 1 h, b) 5 h and c) 1 day.
E.M. Córdoba et al. / Hydrometallurgy 93 (2008) 88–96 93

concluded that covellite forms either aerobically or anaerobically


through the reaction:

CuFeS2 þ Cu2þ Y2CuS þ Fe2þ : ð12Þ

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 previously-
formed elemental sulphur, as follows:

3Cu2þ þ 4H2 O þ 4S-Y3CuS þ HSO þ


4 þ 7H : ð13Þ

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 MacDo-
nald, 1974).
Therefore, chemical reactions (8) and (12) become kinetically more
favorable.

3.2. Influence of iron concentration

The results reported in the previous section showed that


precipitation and nucleation of jarosites on mineral particles led to
passivation of chalcopyrite. In a first 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)
confirmed 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
Fig. 9. X-ray diffractograms of the leaching residues at 68 °C, [Fe]Total = 5 g/L, related to chalcopyrite dissolution is Fe3+. The role of Fe2+, then, would
Einitial = 400 mV and short times: a) 1 h, b) 5 h and c) 1 day.
be to achieve rapid equilibrium of the Fe3+/Fe2+ couple in solution,
thus controlling hydrolysis of the ferric ion, which could be ultimately
The presence of CuS in the leaching products at short times responsible for the passivation of chalcopyrite.
suggests that chalcopyrite dissolves in two steps. First it oxidizes, The positive effect observed in chalcopyrite leaching at 68 °C when
forming CuS as an intermediary product: 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
CuFeS2 þ 2Fe3þ YCuS þ 3Fe2þ þ S-: ð8Þ diffusion of ions towards the chalcopyrite surface. That is consistent
Then, covellite is oxidized by ferric sulphate, releasing Cu2+ ions: with the findings of Hirato et al. (1987) that the dissolution rate at
70 °C increases with ferric ion concentration up to 0.1 M, while above
CuS þ 2Fe3þ YCu2þ þ 2Fe2þ þ S- ð9Þ that value the enhancement is negligible.

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
Table 3
formation of copper-deficient intermediate products, as reported by
EDS microanalysis of chalcopyrite leaching residues at 68 °C, [Fe]Total = 5 g/L,
Ferron (2003) in a study on the leaching of secondary copper minerals: Einitial = 400 mV and short times (results in weight percent)

Cu2 S þ Fe2 ðSO4 Þ3 YCuSO4 þ 2FeSO4 þ CuS ð10Þ Time Chalcopyrite surface S Fe Cu O K P
Original 30.80 32.24 36.96
1h Attacked 30.16 31.99 36.33 1.52
Cu2 S Y Cu1;93–1;96 S Y Cu1;80 S Y CuS : ð11Þ Precipitate 34.06 8.87 48.76 8.31
chalcocite djujerite digenite covellite 5h Attacked 37.32 29.64 33.04
Precipitate 28.99 28.76 30.38 11.87
Some researchers have detected the formation of covellite from 1 day Attacked 32.84 29.74 33.05 4.37
Precipitate 12.39 23.91 4.45 54.68 3.92 0.65
chalcopyrite under reducing conditions. Jang and Wadsworth, 1993,
94 E.M. Córdoba et al. / Hydrometallurgy 93 (2008) 88–96

Fig. 12. X-ray diffractogram 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.
Fig. 10. Influence of iron concentration on the chalcopyrite leaching at 68 °C. 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:
3.3. Influence of temperature
k ¼ A  eEa =RT ð15Þ
The effect of temperature was very pronounced, with negligible
copper dissolution at 35 °C after 13 days of leaching (b3%) vs. almost where Ea is the activation energy, A the pre-exponential factor with
complete mineral dissolution at 68 °C (Figs. 2 and 5). Thermal the same units as the kinetic constant k, R the universal gas constant
activation therefore plays an important role in this process. (8.314 J.K− 1.mol− 1) and T the absolute temperature (K). Then, the
The effect of temperature in the leaching of chalcopyrite with ferric logarithm of the kinetic constant is a linear function of the inverse of
sulphate was evaluated by calculating the activation energy in the temperature. Table 4 shows the values represented in the Arrhenius
experimental temperature range of 35 °C to 68 °C. Four assays were plot in Fig. 15.
performed at 35, 46, 57 and 68 °C, while all other experimental The activation energy for chalcopyrite dissolution, deduced from
conditions were constant, i.e.: 0.5% pulp density, Einitial = 400 mV and the slope of the straight line of the Arrhenius plot, was appreciably
[Fe]Total = 5 g/L. higher (130.7 kJ/mol) than that reported by other researchers (71–
The activation energy was calculated from the kinetic constants 88 kJ/mol) in the range of temperature between 50 and 94 °C and in
and the semilogarithmic plot of the Arrhenius equation. sulphate medium (Dutrizac et al. (1969), Muñoz et al. (1979), Hirato
First, the fraction of reacted chalcopyrite was plotted vs. time for et al. (1987). Those differences could be attributed to the temperature
each temperature (Fig. 13). Then, linear curves were plotted using the range assayed. The kinetic curves show that chalcopyrite dissolution is
simplified shrinking core model proposed by Sohn and Wadsworth negligible at 35 °C, and that the energy barrier responsible for the slow
(1979), according to the following expression: copper dissolution rate could be overcome at 68 °C. However, the
enhancement of copper extraction is less pronounced above a certain
1  ð1  X Þ1=3 ¼ k  t ð14Þ temperature (N50 °C).
Based on the parameters established by Moore (1990) to elucidate
where X is the fraction of reacted chalcopyrite, k the kinetic constant reaction mechanisms in different chemical processes, the high
and t time. That equation is represented graphically in Fig. 14. The activation energy value registered in this study (130.7 kJ/mol)
slope of the final experimental data for the kinetic curves at 57 °C and indicates that during the first stage of chalcopyrite leaching, when
the kinetics is approximately linear, the system is under chemical
control. Therefore, the main drawback of chalcopyrite dissolution

Fig. 11. SEM micrograph of the leaching residue at 68 °C, Einitial = 600 mV and
[Fe]Total = 0.5 g/L. Fig. 13. Influence of temperature on the chalcopyrite leaching at Einitial = 400 mV.
E.M. Córdoba et al. / Hydrometallurgy 93 (2008) 88–96 95

Fig. 15. Arrhenius plot.

1/3
Fig. 14. Variation of 1 − (1 − X) over time.
On the other hand, minority iron-bearing minerals in the starting
mineral (siderite and pyrite) could have an important role in leaching
would be the strong chemical bonding of the crystal lattice, which kinetics. Siderite is quickly dissolved in acidic media, increasing the
requires a high input of energy to the system. These findings support iron concentration in solution:
the hypothesis of Hiskey (1993) that the transport of electrons
through vacants is very poor due to the n-type semiconductivity of FeCO3 þ H2 SO4 YFeSO4 þ CO2 þ H2 O: ð16Þ
chalcopyrite, and hence the first step during chalcopyrite oxidation is
the consumption of vacants to favor electron transportation through In the case of pyrite, there are evidences that this sulphide only
the crystal lattice. Thus, a great deal of the energy applied to the dissolves when chalcopyrite is already passivated. Moreover, pyrite
system in the form of heat is consumed by the displacement of ions rest potential is higher than that of chalcopyrite and, therefore, pyrite
from the bulk of the particle to the surface, eliminating surface vacants enhances dissolution chalcopyrite through galvanic contact.
and favoring the transport of electrons through the chalcopyrite Finally, the comparison between studies on chalcopyrite leaching
surface. is a difficult task because of differences in experimental conditions
At the same time, increasing iron concentration from 0.5 to 5 g/L used. However, unlike at high temperature, chalcopyrite dissolution
improves the leaching rate at 68 °C (Fig. 10), and so the diffusion of and ferric ion hydrolysis kinetics are very slow processes at low
ferric ion to the chalcopyrite surface may also control the process. temperature (35 °C).
Researchers have not reached a consensus on the control steps
governing chalcopyrite leaching. Our findings are similar to those 4. Conclusions
reported by Linge (1976) and Hirato et al. (1987) with a linear branch
in the copper kinetic curve. Nevertheless, many researchers (Dutrizac 1. The redox potential is a key factor in the leaching of chalcopyrite. A
et al., 1969; Ferreira and Burkin, 1975; Muñoz et al., 1979; Parker et al., high potential at the onset of leaching provokes rapid passivation of
1981; Kametani and Aoki, 1985; Hackl et al., 1995, etc.) have obtained chalcopyrite.
parabolic kinetics with an initial linear branch, similar to our results 2. Ferric/ferrous sulphate leaching solutions tend to reach equilibrium
but with very varied conclusions. This variety in the hypotheses to be when the activities of the two ions are equal, which is associated
found in the literature reflects a lack of agreement on the mechanisms with a critical potential of approximately 450 mV. When the redox
of chalcopyrite dissolution. potential is very high initially, that tendency to equilibrium favors
These results seem to indicate that hydrolysis and precipitation of rapid precipitation of ferric ion as jarosite and consequently
Fe3+ has a main role in chalcopyrite passivation in ferric solutions by passivation of chalcopyrite.
preventing the contact between the mineral surface ant the oxidizing 3. The activation energy during chalcopyrite leaching was 130.7 kJ/
agent in solution. mol, which is a clear demonstration of the importance of thermal
Decreasing pH or removing some ions that favor iron precipitation activation in this process.
could be a way to prevent ferric ion hydrolysis. However, our attempts 4. Increasing the iron concentration from 0.5 to 5 g/L had a positive
in that direction (results not shown) have not solved the problem of effect in the chalcopyrite leaching at 68 °C.
chalcopyrite passivation. Copper dissolution rate decreased when pH 5. Chalcopyrite dissolves through the intermediate formation of
was reduced from 0.5 to 2.0, perhaps because of that the species covellite, CuS, which is later oxidized by ferric ion to release Cu2+
responsible for the oxidation of chalcopyrite is not properly Fe3+ but ions:
probably Fe(SO4)−2 (Córdoba, 2005). Furthermore, the removal of
monovalent cations (K+, Na+ or NH+4) from solution neither prevents CuFeS2 þ 2Fe3þ YCuS þ 3Fe2þ þ S-
hydrolysis and precipitation of iron as goethite and hydronium
jarosite that, like potassium jarosite, also tend to nucleate over
chalcopyrite particles. CuS þ 2Fe3þ YCu2þ þ 2Fe2þ þ S-:

6. The elemental sulphur that forms during chalcopyrite leaching is


porous and does not form a passivating layer on the chalcopyrite
Table 4 surface.
Values represented in the Arrhenius plot of Fig. 15

T (K) k (days− 1) k (s− 1) 1000 / T (K− 1) ln k (s− 1) References


−9
308 0.0006 6.9444 ⁎ 10 3.25 −18.79
319 0.0044 5.0926 ⁎ 10− 8 3.13 −16.79 Ammou-Chokroum, M., Cambazoglu, M., Steinmez, D., 1977. Oxydation menagée de la
330 0.0319 3.6921 ⁎ 10− 7 3.03 −14.81 chalcopyrite en solution acide: analyses cinétique de réactions. II. Modéles
diffusionales. Bulletin de la SocieÂte francÐaise de mineÂralogie et de
341 00732 8.4722 ⁎ 10− 7 2.93 −13.98
cristallographie 100, 161–177.
96 E.M. Córdoba et al. / Hydrometallurgy 93 (2008) 88–96

Baron, D., Palmer, C.D., 1996. Solubility of jarosite at 4–35 °C. Geochimica et Hiskey, J.B., 1993. Chalcopyrite semiconductor electrochemistry and dissolution. In:
Cosmochimica Acta 60 (2), 185–195. Reddy, R.G., Weizenbach, R.N. (Eds.), The Paul E. Queneau International Symposium,
Barr, D.W., Jordan, M.A., Norris, P.R., Phillips, C.V., 1992. An investigation into bacterial Extractive Metallurgy of Copper, Nickel and Cobalt. Volume I: Fundamental Aspects.
cell, ferrous iron, pH and Eh interactions during thermophilic leaching of copper The Minerals, Metals & Materials Society, pp. 949–969.
concentrates. Minerals Engineering 5 (3–5), 557–567. Jang, J.H., Wadsworth, M.E., 1993. Hydrothermal conversion of chalcopyrite under
Bigham, J.M., Schwertmann, U., Traina, S.J., Winland, R.L., Wolf, M., 1996. Schwertman- controlled Eh and pH. In: Reddy, R.G., Weizenbach, R.N. (Eds.), The Paul E.
nite and the chemical modeling of iron in acid sulfate waters. Geochimica et Queneau International Symposium, Extractive Metallurgy of Copper, Nickel and
Cosmochimica Acta 60 (12), 2111–2121. Cobalt. Volume I: Fundamental Aspects. The Minerals, Metals & Materials Society,
Burkin, A.R., 1969. Solid-state transformations during leaching. Minerals science and pp. 689–707.
engineering 1 (1), 4–14. Kametani, H., Aoki, A., 1985. Effect of suspension potential on the oxidation rate of
Cabral, T., Ignatiadis, I., 2001. Mechanistic study of the pyrite–solution interface during copper concentrate in a sulfuric acid solution. Metallurgical Transactions B 16B,
the oxidative bacterial dissolution of pyrite (FeS2) by using electrochemical 695–705.
techniques. International journal of mineral processing 62, 41–64. Levenspiel, O., 2004. Capítulo 3: Interpretación de los datos obtenidos en un reactor
Casas, J.M., Lienqueo, M.E., Cubillos, F., Herrera, L., 2000. Modelación cinética de la intermitente, Ingeniería de las reacciones químicas, 3a edición. Editorial Limusa
precipitación de hierro como jarosita en soluciones lixiviantes utilizando la bacteria Wiley, México, pp. 38–67.
Thiobacillus ferrooxidans. Congreso Chileno de Ingeniería Química. Universidad de Linge, H.G., 1976. A study of chalcopyrite dissolution in acid ferric nitrate by
Santiago. Octubre de 2000. potentiometric titration. Hydrometallurgy 2, 51–64.
Córdoba, E.M. (2005). Nuevas evidencias sobre los mecanismos de lixiviación y Logan, S.R., 2000. Capítulo 1: Fundamentos empíricos de la cinética química. Funda-
biológica de la calcopirita. Tesis Doctoral. Departamento de Ciencia de los Materiales mentos de cinética y química. Editorial Addison Wesley, Madrid, pp. 1–22.
e Ingeniería Metalúrgica. Facultad de Ciencias Químicas. Universidad Complutense Lowson, R., 1982. Aqueous oxidation of pyrite by molecular oxygen. Chemical Reviews
de Madrid. 82 (5), 461–497.
Dry, M.J., Bryson, A.W., 1988. Prediction of redox potential in concentrated iron sulphate Majima, H., Awakura, Y., Hirato, T., Tanaka, T., 1985. The leaching of chalcopyrite in ferric
solutions. Hydrometallurgy 21, 59–72. chloride and ferric sulfate solutions. Canadian Metallurgical Quarterly 24 (4),
Dutrizac, J.E., 1978. The kinetics of dissolution of chalcopyrite in ferric ion media. 283–291.
Metallurgical Transactions B 9B, 431–439. Moore, J.J. (1990). Chapter 3: Reaction kinetics. Chemical Metallurgy. Ed. Butterworth
Dutrizac, J.E., 1983. Factors affecting alkali jarosite precipitation. Metallurgical Heinemann. 2nd edition, Great Britain, 95–126.
Transactions B 14B, 531–539. Muñoz, P.B., Miller, J.D., Wadsworth, M.E., 1979. Reaction mechanism for the acid ferric
Dutrizac, J.E., 1989. Elemental sulphur formation during the ferric sulphate leaching of sulfate leaching of chalcopyrite. Metallurgical Transaction B 10B, 149–158.
chalcopyrite. Canadian Metallurgical Quarterly 28 (4), 337–344. Nicol, M.J., Lázaro, I., 2003. The role of non-oxidative processes in the leaching of
Dutrizac, J.E., MacDonald, R.J.C., 1974. Ferric ion as a leaching medium. Minerals science chalcopyrite. In: Riveros, P.A., Dixon, D., Dreisinger, D.B., Menacho, J. (Eds.), COPPER
and engineering 6 (2), 59–100. 2003. Volume VI—Hydrometallurgy of Copper (Book 1), pp. 367–381. Santiago,
Dutrizac, J.E., MacDonald, R.J.C., Ingraham, T.R., 1969. The kinetics of dissolution of Chile.
synthetic chalcopyrite in aqueous acidic ferric sulfate solutions. Transactions of the Norris, P.R., Barr, D.W., 1985. Growth and iron oxidation by acidophilic moderate
Metallurgical Society of AIME 245, 955–959. thermophiles. FEMS microbiology letters 28, 221–224.
Ferreira, R.C.H., Burkin, A.R., 1975. Acid leaching of chalcopyrite. In: Bulkies, A.R. (Ed.), Okamoto, H., Nakayama, R., Tunekawa, M., Hiroyoshi, N., 2003. Improvement of
Leaching and Reduction in Hydrometallurgy. Inst. Min. Met. , pp. 54–56. Londres. chalcopyrite leaching in acidic sulfate solutions by redox potential control. In:
Ferron, C.J., 2003. Leaching of secondary copper minerals using regenerated ferric Riveros, P.A., Dixon, D., Dresinger, D.B., Menacho, J. (Eds.), COPPER 2003. Volume VI—
sulphate. In: Riveros, P.A., Dixon, D., Dreisinger, D.B., Menacho, J. (Eds.), COPPER Hydrometallurgy of Copper (Book 1), pp. 67–81. Santiago, Chile.
2003, Vol. VI — Hydrometallurgy of Copper (Book 1), pp. 337–352. Santiago, Chile. Parker, A.J., Paul, R.L., Power, G.P., 1981. Electrochemical aspects of leaching copper from
Garrels, R.M., Christ, C.L., 1965. Solution, Minerals, and Equilibria. Editorial Harper & chalcopyrite in ferric and cupric salt solutions. Australian journal of chemistry 34,
Row, New York, pp. 213–233. 13–34.
Hackl, R.P., Dreisinger, D.B., Peters, E., King, J.A., 1995. Passivation of chalcopyrite during Silverman, M.P., Lundgren, D.G., 1959. Studies on the chemoautotrophic iron bacterium
oxidative leaching in sulfate media. Hydrometallurgy 39, 25–48. Ferrobacillus ferrooxidans. An improved medium and a harvesting procedure for
Hirato, T., Majima, H., Awakura, Y., 1987. The leaching of chalcopyrite with ferric sulfate. securing high cell yields. Journal of bacteriology 77, 642.
Metallurgical Transactions B 18B, 489–496. Sohn, H.Y., Wadsworth, M.E., 1979. Rate Processes of Extractive Metallurgy. Plenum,
Hiroyoshi, N., Hirota, M., Hirajima, T., Tsunekawa, M., 1997. A case of ferrous sulfate New York, p. 133.
addition enhancing chalcopyrite leaching. Hydrometallurgy 47, 37–45. Third, K.A., Cord-Ruwisch, R., Watling, H.R., 2000. The role of iron-oxidizing bacteria in
Hiroyoshi, N., Hirota, M., Hirajama, T., Tsunekawa, M., 1999. Inhibitory effect of iron- stimulation or inhibition of chalcopyrite bioleaching. Hydrometallurgy 57,
oxidizing bacteria on ferrous-promoted chalcopyrite leaching. Biotechnology and 225–233.
Bioengineering 64 (4), 478–483. Third, K.A., Cord-Ruwisch, R., Watling, H.R., 2002. Control of the redox potential by
Hiroyoshi, N., Miki, H., Hirajima, T., Tsunekawa, M., 2000. A model for ferrous-promoted oxygen limitation improves bacterial leaching of chalcopyrite. Biotechnology and
chalcopyrite leaching. Hydrometallurgy 57, 31–38. Bioengineering 78 (4), 433–441.
Hiroyoshi, N., Miki, H., Hirajima, T., Tsunekawa, M., 2001. Enhancement of chalcopyrite
leaching by ferrous ions in acidic ferric sulfate solutions. Hydrometallurgy 60,
185–197.

Anda mungkin juga menyukai