Sunteți pe pagina 1din 18

Hydrometallurgy 140 (2013) 163180

Contents lists available at ScienceDirect

Hydrometallurgy
journal homepage: www.elsevier.com/locate/hydromet

Review

Chalcopyrite hydrometallurgy at atmospheric pressure: 1. Review of


acidic sulfate, sulfatechloride and sulfatenitrate process options
H.R. Watling
CSIRO Minerals Down Under, CSIRO Process Science and Engineering, P.O. Box 7229, Karawara, WA 6152, Australia

a r t i c l e i n f o a b s t r a c t

Article history: The need to process low-grade and/or complex chalcopyrite-containing ores that cannot be concentrated is the
Received 13 May 2013 main driver for the development of hydrometallurgical processes. The ferric sulfatesulfuric acid system, with or
Received in revised form 19 August 2013 without the assistance of microorganisms, has been studied extensively because it comprises the most promising,
Accepted 22 September 2013
low-cost process route. Alternative oxidants to ferric ion are known but, as yet, their superior oxidation strengths
Available online 10 October 2013
have not been exploited other than at laboratory scale, probably due to their higher costs. Hybrid sulfatechloride
Keywords:
and sulfatenitrate systems were included because they may offer specic advantages in some instances. The
Chalcopyrite aims of this review were to summarise current knowledge in respect of these systems and highlight potentially
Leaching rewarding areas for future research.
Dissolution 2013 Published by Elsevier B.V.
Passivation
Suldes

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164
1.1. Process options for chalcopyrite concentrates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164
1.2. Process options for low-grade chalcopyrite ores . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
1.3. Scope of this review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
2. Sulfuric acidferric sulfate systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
2.1. Chemistry of leaching . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
2.2. Chalcopyrite surface overlayers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
2.3. Chalcopyrite surface structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
2.4. Redox control in ferric sulfate systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
2.5. Microorganisms as catalysts in ferric sulfate systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
2.5.1. Ambient- to moderate-temperature bioleaching . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
2.5.2. Bioleaching at moderate to high temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
2.5.3. Reduced sulfur additives to increase extraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
2.5.4. Separation of biological and chemical processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
2.6. Cations as catalysts in ferric sulfate systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
3. Sulfuric acid alternative oxidants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172
3.1. Sulfuric aciddichromate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172
3.2. Sulfuric acidchlorate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
3.3. Sulfuric acidpermanganate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
3.4. Sulfuric acidhydrogen peroxidehydroxyl radical . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
3.5. Sulfuric acidozone . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
3.6. Sulfuric acidperoxodisulfate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
4. Hybrid sulfatechloride systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
4.1. H2SO4NaClO2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
4.2. H2SO4Fe2(SO4)3NaCl. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
4.3. H2SO4Fe2(SO4)3LiCl . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176

E-mail address: Helen.Watling@csiro.au.

0304-386X/$ see front matter 2013 Published by Elsevier B.V.


http://dx.doi.org/10.1016/j.hydromet.2013.09.013
164 H.R. Watling / Hydrometallurgy 140 (2013) 163180

5. Hybrid sulfatenitrate systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176


5.1. Sulfuric acidnitric acid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
5.2. Sulfuric acidsodium nitrate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
6. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178

1. Introduction low-grade oxide and secondary suldes, it is clear that the proportion
of low-grade ores containing chalcopyrite will increase in the future.
There is currently an imbalance between copper supply and world Therefore, there are imperatives to improve hydrometallurgical tech-
demand (Fig. 1). The lack of new, large, high-grade deposits to replace nologies for the extraction of copper from chalcopyrite in ores of such
those that are nearing the ends of their lives, together with the low grades that they are uneconomic to concentrate, and to augment
increased delays between discovery and production due to higher the supply by extracting copper from polymetallic ores of complex
industry standards and extensive permitting requirements, means mineralogy that cannot be concentrated and from chalcopyrite con-
that the current global imbalance between supply and demand could centrates that contain undesirable impurities such as arsenic and cannot
continue for several years. be smelted.
At the same time, there has been a decline in copper grades, often
remarked upon as a future challenge to the copper industry. As an
example, at Escondida Mine, Chile, the world's largest copper mine 1.1. Process options for chalcopyrite concentrates
and producer of 9.5% of the global copper supply, the average copper
content of mined ore fell from 1.65% in 20072008 to 1.14% in In recent years, many processes for the extraction of copper from
20112012, with measured resource, indicated resource and inferred chalcopyrite concentrates have been developed, some of them at
resource all substantially less than 1% grade (Basto, 2012). In his review atmospheric pressure. In these processes, high copper extractions
of historical trends in Australian mining, Mudd (2010) presented data could be achieved, soluble copper could be separated and puried
for copper grades mined in the last 150 years and concluded that using well-established technologies such as solvent extraction and
the decline in ore grades would continue and, in addition, that ore electrowinning, and pure, high-quality metal products were recov-
mineralogy would become more complex and make the low-grade erable (Dreisinger, 2006). Using examples from acidic sulfate-based
ores more difcult to process. Strategies to reduce the imbalance chemical systems, four strategies employed to maximise copper ex-
between supply and demand could include the processing of complex traction particularly from chalcopyrite were: (i) as high a temperature
ores, the recycling of metals from electronic and other copper- process as could be reasonably managed at atmospheric pressure (e.g.,
containing waste materials and the development of processes to extract BioCOP Batty and Rorke, 2006); (ii) ne grinding to increase chal-
copper from dirty concentrates containing penalty elements or from copyrite reactivity and overcome passivation (e.g., BacTech/Mintek
low grade ores and mine waste materials. van Staden, 1998); (iii) the use of additives (e.g., Galvanox Dixon
These strategies are not new. Heap, dump, in situ and vat leaching of et al., 2008); and (iv) exploitation of innovative combinations of
whole ores are the preferred technologies for the processing of low- processing technologies, such as heap leaching of concentrates (e.g., the
grade ores. Data from 2010 showed that proven concentration and Geocoat process Harvey and Bath, 2007).
pyrometallurgical technologies accounted for 80% of world copper Most of the processes were taken to pilot scale and some were dem-
production while hydrometallurgical processing of low-grade ores onstrated at a larger scale. However, there are still few commercial-
containing copper oxides or secondary copper suldes contributed scale operations and energy-intensive processes are unlikely to be
about 20% of annual copper production (Index Mundi, 2013). As economic for concentrates, except in circumstances where the
chalcopyrite (CuFeS2) is the most abundant but also the most refractory competing pyrometallurgical technologies cannot be employed. Based
of the copper suldes, and with the current extensive exploitation of on a survey of process options for copper concentrates assisted by the
application of a qualitative ranking technique, Lunt et al. (1997)
concluded that two of the best options for the hydrometallurgical
processing of concentrates were sulfationroastleach and bioleach.
6
They also noted that pyrometallurgical process routes offered some
distinct advantages, not least low cost and industry acceptance and
copper production surplus or
deficit [as % of production]

4
that for any given project there were a limited number of suitable
process options. Similar points were made by Peacey et al. (2004),
2 who outlined a potential opportunity as being the hydrometallurgical
treatment of dirty copper concentrates with subsequent discharge of
0 the acidic leachate to a heap leach/SX/EW operation in circumstances
where copper concentrate was readily available but ore for heap
-2 leaching was diminishing.
Dreisinger (2006) summarised nine reasons why hydromet-
-4
allurgical processes for the treatment of copper sulde concentrates
had failed to achieve sustained commercial production. They can be
distilled into: (i) Production: incomplete copper and precious metal
-6
recovery and/or poor product quality; (ii) Wastes: difculties with
2011
[2012]
[2013]
2006

2010
2000
2001
2002
2003
2004
2005

2007
2008
2009

treatment and/or disposal of insoluble residues with particular


emphasis on sulfur and (iii) Technoeconomics: costs not competitive
Fig. 1. Imbalance between copper production and copper demand. Annual surpluses or
with pyrometallurgy and higher perceived risks associated with
decits (data from International Copper Study Group, 2012). Square brackets indicate implementing new processes. Peacey et al. (2004) and Dreisinger
predictions. (2006) noted that successful processes tended to represent niche
H.R. Watling / Hydrometallurgy 140 (2013) 163180 165

90
opportunities and predicted that the eld would continue to advance by Djurleite
80
way of necessity or unique opportunity. Other criteria for successful Bornite

Cu extracted (%)
processes could include (iv) Universality: applicable to all copper 70
concentrates and able to deal with their various impurities, for example, 60 Covellite
Zn, Pb, As, Sb, Bi, Se, Hg and F and (v) Simplicity: robust, reliable 50
pyritic chalcopyrite
technology suitable for remote locations (Jones, 1996). 40
30 porphyry chalcopyrite
20 chalcocite (+digenite)
1.2. Process options for low-grade chalcopyrite ores 10
0 chalcopyrite (+ 1500
0 5 10 15 20 25 30 35 g/t Ag)
Heap and dump technologies were developed to overcome the
Time (days)
challenges of low copper grade in vast quantities of ores. These
technologies may also be applied to less-difcult ores, not necessarily Fig. 2. Copper extraction from copper sulde minerals (bioleaching, pH 1.82, 30 C, 2 wt.%
of the lowest grade, for (i) smaller-sized ore deposits, (ii) deposits in mineral).
remote locations lacking the necessary infrastructure for, or access to, Data from Fu et al. (2012), Johnson et al. (2008), and Ruan et al. (2010).
pyrometallurgical process routes, or (iii) when pyrometallurgical
processing is marginal or uneconomic. Copper recovery rates during
leaching in acidic systems at atmospheric pressure and ambient Heap leaching of copper oxide ores and bioleaching of secondary
temperature vary greatly for different mineral phases (e.g., Table 1). copper sulde (chalcocite) ores is widely practised (Domic, 2007).
Rates are inuenced by one or more of the physico-chemical properties However, heap leaching of chalcopyrite has yet to be implemented at
of the ores: commercial scale and the extraction of copper from dumps of low-
grade chalcopyrite ores or tailings is a practical option only because
The copper minerals present (Fig. 2). As a general rule M2S is more the slow and low copper recoveries are proportionate to the low
readily dissolved than MS and impurities also inuence dissolution processing costs (Schnell, 1997). In their reviews, Watling (2006) and
rates. Pradhan et al. (2008) focused on the potential of heap leaching as a
The mineral associations and copper mineral liberation (Fig. 3). process route for low-grade chalcopyrite ores. However, among current
For the example ore, QEMSCAN analysis indicated that only the and past heap or dump leaching operations, only one chalcopyrite heap
0.85 mm size fraction contained substantial liberated chalcopyrite; leach operation (Straits Resources' Girilambone Copper Company,
b30% of the chalcopyrite in the larger size fractions had surface NSW) was identied and two large-scale test heaps of chalcopyrite
expression and, overall, only about 40% of the ore copper content ore were described or discussed (Schlitt, 2006; van Staden et al., 2005;
could be leached. Particle structure analysis using X ray tomog- Watling, 2006). Watling (2006) expressed the opinion that the efcient
raphy failed to reveal particle fracturing that might have assisted in heap leaching of chalcopyrite would require greater management and
exposing further chalcopyrite grains to the leachate. control than was thus far required in oxide and secondary sulde
The chalcopyrite grain size. For concentrates, the need to ne grind heap leaching and van Staden et al. (2005) described some of those
chalcopyrite concentrates to obtain a reasonable (economic) copper
extraction rate during leaching is widely reported and constitutes a
key parameter in the Albion process (Hourn and Halbe, 1999; Hourn
et al., 1999) and the Mintek/BacTech process (Gericke et al., 2009;
van Staden, 1998; Wang, 2005). The same effect applies to large- Feldspar
grained chalcopyrite in ores, assuming that the grains are exposed
to the leachate (e.g., Naderi et al., 2011). K-Feldspar
Gangue mineral dissolution consumes acid during leaching or Titanite
bioleaching and may form amorphous silica gel, thus increasing the
viscosity of leachates. Gangue mineral dissolution may increase iron
concentrations in leachates and therefore promote jarosite formation, Quartz
and may release potentially toxic elements to bioleaching solutions
(Dopson et al., 2009; Halinen et al., 2009; Watling et al., 2009).
Chalcopyrite
Table 1 Pyrite
Comparative dissolution rates in laboratory-scale tests using acidic sulfate systems for
different copper-bearing minerals found in heaps and dumps of low-grade ores.
Biotite
Required Mineral Ideal formula

Hours to days Atacamite Cu2Cl(OH)3


Chrysocolla CuSiO32H2O
Neotocite (Cu,Mn)2H2Si2O5(OH)4nH2O
Tenorite CuO
Malachite Cu2(CO3)(OH)2
Quartz
Azurite Cu3(CO3)2(OH)2
Antlerite CuSO4.2Cu(OH)2
K-Feldspar
Brochantite CuSO4.3Cu(OH)2
Days to months Native Cu Cu
Cuprite Cu2O
Chalcocite Cu2S
Months to years Bornite Cu5FeS4
Covellite CuS
Fig. 3. Ore particle (N7 mm, 0.7% Cu). QEMSCAN mineral association analysis of multiple
Enargite Cu3AsS4
blocks of each size fraction of this ore revealed that chalcopyrite was primarily associated
Chalcopyrite CuFeS2
with pyrite and quartz/feldspar.
166 H.R. Watling / Hydrometallurgy 140 (2013) 163180

requirements and how they were being achieved in a pilot heap of but rather the aim was to inform researchers, metallurgists and plant
chalcopyrite ore. Vat leaching, largely superseded by heap leaching, is operators about the wide variety of chemical systems that might be
applied at Mantos Blancos, Chile (Schlesinger et al., 2011) for the applied in the future when copper demand is higher, ore grades
extraction of copper from rapid-leaching minerals such as copper are lower and new technologies have been developed. While the
oxides or carbonates and may enjoy a revival and wider application advantages or disadvantages of current technologies might be referred
with the development of continuous vat technologies (Mackie and to in the contexts of reported results or applications of specic systems,
Trask, 2009; Schlitt and Johnston, 2010). a detailed account of the engineering of such technologies, their
The Geocoat bioleaching technology is a hybrid technology. It management and/or control are outside the scope of the review.
combines the high recoveries associated with reactor leaching of
chalcopyrite concentrate with the lower capital cost of heap leaching. 2. Sulfuric acidferric sulfate systems
In this process, copper concentrate is coated as a thickened slurry onto
host rock particles (625 mm). The coated host rock particles are The most commonly employed hydrometallurgical process for the
then stacked and irrigated as for heap leaching (Harvey and Bath, oxidation of chalcopyrite and extraction of copper is the sulfuric acid
2007). Sulde dissolution is catalysed by acidophilic microorganisms ferric sulfate system. It is the system of choice for bioleaching processes
appropriate to the heap temperature. The complementary Geoleach at atmospheric pressure, including stirred tank technologies for
technology comprised a control strategy of aeration and irrigation concentrates (Batty and Rorke, 2006) and heap or dump technologies
rates to maximise microbial activity and heat generation and/or for low-grade ores (Watling, 2006). However, whether chemical or
conservation within heaps, intended to promote faster leaching, bio-assisted chemical leaching is undertaken, the oxidation of chal-
particularly of chalcopyrite. Though widely tested and publicised, the copyrite is slow and incomplete, possibly the consequence of the
technology has not been commercialised. mineral crystalline structure and changes therein (de Oliveira et al.,
2012; Klauber, 2003) but also inhibited by insoluble secondary reaction
1.3. Scope of this review products (elemental sulfur, ferric hydroxides or hydroxysulfates)
forming what are termed passivation or overlayers on chalcopyrite
The purposes of this review were to describe the chemistry of surfaces. There is value in examining the chemistry of ferric sulfate
chalcopyrite leaching at atmospheric pressure, in sulfate media with and acid leaching systems, some of which are summarised in Table 2,
different oxidants or reductants and/or other additives and treatments because they comprise the basic chemical systems against which the
and, where possible, to compare copper extractions. Not surprisingly, use of other chemical systems would be assessed during the
most research on chalcopyrite leaching has been undertaken using development of alternative atmospheric-pressure leaching processes.
chalcopyrite concentrates, whether or not the target application was
the processing of concentrate or of low-grade ore. Such tests utilised a 2.1. Chemistry of leaching
simple matrix, largely free of gangue minerals that might obscure
important relationships between starting and product materials. The generally accepted reactions for the extraction of copper from
However, one of the difculties of making direct comparisons between chalcopyrite via oxidation (oxygen, ferric ions) or acid (H2SO4) leaching
the results of published studies has been the failure of many researchers are:
to include sufcient detail in methods, or test-mineral properties.
Where data are presented in different formats, or disguised for reasons CuFeS2 2Fe2 SO4 3 CuSO4 5FeSO4 2S
0
1
of condentiality, comparisons between studies become even more
difcult. 0
CuFeS2 O2 2H2 SO4 CuSO4 FeSO4 2S 2H2 O 2
The review was limited to those sulfate-based systems operated at
ambient pressure within the temperature range constrained by that
condition. Chloride, nitrate and other acidic leaching systems, and pre- 4FeSO4 2H2 SO4 O2 2Fe2 SO4 3 2H2 O: 3
and co-treatments associated with chalcopyrite hydrometallurgy will
be reviewed separately. Sulfate systems operating at above ambient Dutrizac (1981) reviewed the literature on ferric ion leaching of
pressure and higher temperatures were comprehensively described chalcopyrite (reaction (1)) and, with data from ancillary studies,
and discussed by Dreisinger (2006) and McDonald and Muir (2007a, summarised existing knowledge at that time as: (i) leaching was
b) and references therein, and are not discussed further. In the present independent of acid above that required to keep iron in solution (and
review, no account was taken of the possible economics of processing, iron need not exceed 0.1M Fe2(SO4)3 Parker et al., 1981); (ii) leaching

Table 2
Acidic sulfate processes developed for chalcopyrite or other copper sulde concentrates or ores and operated at atmospheric pressure.

Process Temperature Size Differentiating conditions Scale


(C) (m)

Heap leaching (e.g., Quebrada Blanca, Chile) (Domic, 2007) Ambient Crushed ore, stacked, irrigated with dilute H2SO4 solution C
ROM dump leaching (e.g., La Escondida, Chile) (Domic, 2007) Ambient ROM ore stacked and irrigated with dilute H2SO4 C
Geocoat (Harvey and Bath, 2007) Ambient n.d. Sulde concentrate supported on host rock particles; sulfate D
bioleaching in heaps
BacTech/Mintek (Gericke et al., 2009; van Staden, 1998; Wang, 2005) 3550 510 Sulfate bioleach; ne grind D
BioCOP (Batty and Rorke, 2006) 6580 37 Sulfate leach at higher temperature using archaea as catalysts D
Sepona (Baxter et al., 2003) 80 100 Ferric sulfate leach of chalcocite rimming pyrite (FeS2) C
Galvanox (Dixon et al., 2008) 80 5375 Ferric sulfate leach; pyrite to chalcopyrite in 2:1 ratio P
Albion (Hourn and Halbe, 1999; Hourn et al., 1999) 85 510 Ferric sulfate leach; nely ground concentrate P
Cobre Las Cruces Project (Fleury et al., 2010) 90 150 Ferric sulfate generated from pyrite oxidation with O2; mainly Cu9S5 C
and Cu2S (6.2 wt.%), minor CuFeS2
Cuprochlor process (Espejo et al., 2001; Herreros et al., 2006) Ambient Sulfatechloride leach with CaCl2 agglomeration C
Nitric acid route (Bjorling et al., 1976) 90 n.d. Sulfatenitric leach of chalcopyrite concentrate L
a
In the Sepon process, residual pyrite is oated and then oxidised under pressure to generate acid and ferric ions for use in the secondary heap leach. C commercial; D demonstration; L
laboratory P pilot; n.d. not disclosed.
H.R. Watling / Hydrometallurgy 140 (2013) 163180 167

rates were proportional to chalcopyrite surface area, only slightly a


dependent on ferric ion concentrations and increased with increased
temperature; and (iii) increased sulfate ion concentration contributed
to slower leaching rates. Hiroyoshi et al. (1997) proposed that
chalcopyrite was oxidised by dissolved oxygen (reaction (2)) and that
the generated ferrous ions were oxidised to ferric ions (reaction (3)).
Their evidence for these reactions was that the pH increased with
time when ferrous sulfate was added to acid medium containing chal-
copyrite, indicating the consumption of protons. Subsequent studies
by the authors were incorporated into a two-stage reaction model to
explain ferrous-promoted chalcopyrite dissolution (Hiroyoshi et al., CuFeS2
2000 and references therein) in which (i) chalcopyrite was reduced
by ferrous ions in the presence of cupric ions to form chalcocite and
(ii) chalcocite was oxidised (more readily than chalcopyrite) by
dissolved oxygen and/or ferric ions to form cupric ions and elemental
sulfur insoluble product. From thermodynamic calculations, chalcocite b
formation only occurred when the redox potential of the solution
was lower than the critical potential (a function of the ferrous and
cupric ion activities); and the optimum oxidationreduction potential
(ORP) for chalcopyrite leaching increased with increased copper(II)
concentrations. Conversion of the normalised redox potential to the
solution redox potential showed the optimum ORP for chalcopyrite
leaching to be a function of cupric and ferrous ion concentrations
(Hiroyoshi et al., 2008). Sandstrm et al. (2005) examined the reaction
products formed during the chemical leaching of chalcopyrite and
found Cu(I) species in all samples, consistent with the Hiroyoshi et al.
(2000) two-stage model. CuFeS2
Nicol and Lzaro (2003) showed that dissolution of chalcopyrite
could occur in the absence of any oxidising reagent at potentials
lower than 0.4 V versus Standard Hydrogen Electrode (SHE), with the
formation of a detectable soluble sulfur species, such as hydrogen Fig. 4. Examples of reaction products on chalcopyrite surfaces: (a) ferric hydroxysulfate
layer after 96 h exposure to ferric sulfate solution (pH 1.8); (b) elemental sulfur after 2 h
sulde. They presented data indicating that reaction (4) was the most
exposure to HOCl (pH 4).
proton-consuming reaction and hypothesised that it was governed by
two steps: (i) rapid dissolution to establish the equilibrium between
soluble species at the chalcopyrite surface and the bulk solid and
In a critical review of the surface science investigations of what
(ii) rate-determining diffusion of the soluble species away from the
might be responsible for hindered chalcopyrite dissolution in ferric
surface. Nicol and Lzaro (2003) examined the thermodynamic
sulfate systems, Klauber (2008) claried some of the discrepancies
feasibility of reaction (4) at 25 C and higher temperatures using
between studies. Klauber (2008) concluded that polysuldes could be
Outukumpu HSC software and the CrissCobble technique.
rejected as passivation candidates because they were too reactive
2 2 and oxidised to elemental sulfur on exposure to air, especially in the
CuFeS2 4H Cu Fe 2H2 S: 4
presence of moisture and metal-decient suldes and that the presence
of metal decient suldes was questionable. Overlayers of elemental
2.2. Chalcopyrite surface overlayers sulfur formed on chalcopyrite surfaces could hinder dissolution but
might subsequently be peeled off, allowing further dissolution. Overlayers
There is a general view that the slow dissolution rates of chalcopyrite of iron(III)-insoluble compounds such as jarosite were an inevitable
in ferric sulfate leaching media are, in part, a consequence of the consequence of long leach times, triggered by the formation of an
formation of secondary reaction product on the mineral surface. These intermediate ferric sulfate compound at the chalcopyrite surface as an
secondary reaction products, passivation or overlayers, may be of integral part of the ferric ion leaching mechanism.
such depth and sufciently compact (Fig. 4) as to hinder the diffusion More recently, Debernardi and Carlesi (2012) presented a review of
of reagents to the chalcopyrite surface and/or the diffusion of ions chalcopyrite passivation in which they outlined the capabilities and
away from the surface, or they may have low electrical conductivity. limitations of electrochemical and surface analysis techniques in
Reaction products thought to be implicated in chalcopyrite passivation resolving the nature of chalcopyrite passivation. As part of their review,
included: in which they summarised many of the papers discussed by Klauber
Complex lms of suldes, polysuldes and/or elemental sulfur (2008), Debernardi and Carlesi (2012) questioned whether the current
(reaction (1)), or an unreactive, copper-rich layer on partially- techniques in electrochemistry and surface analysis could provide
oxidised chalcopyrite surfaces (Fu et al., 2012; Hackl et al., 1995; reliable data on surface phenomena and whether those data were
Linge, 1976; Nicol and Lzaro, 2003; Warren et al., 1982); being interpreted correctly. They identied potential research topics
Iron oxides, hydroxides, iron hydroxysulfates and/or jarosite that might assist in the future interpretation of data obtained using
(reaction ((5)) (Crdoba et al., 2008a,b; Gmez et al., 1996; Parker sophisticated instrumentation.
et al., 2003; Sandstrm et al., 2005; Stott et al., 2001).
2.3. Chalcopyrite surface structure
3 2
3Fe 2SO4 6H2 O M MFe3 SO4 2 OH6 6H 5
In their initial study, using X-ray photoelectron spectroscopy (XPS),
Klauber et al. (2001) proposed that a disulde phase was among
where M = K+, Na+, NH+ +
4 or H . the reaction products on leached chalcopyrite surfaces (Fig. 5). The
168 H.R. Watling / Hydrometallurgy 140 (2013) 163180

Fig. 5. X-ray photoelectron spectrum of a chalcopyrite surface after treatment with 0.1 M ferric sulfate solution at pH 1.8. Deconvolution of the spectrum to reveal the underlying sulde,
and the disulde, elemental sulfur and sulfate species on the surface (Klauber et al., 2001).

synthesis of model polysuldes and their subsequent examination using layer with lower coordination number than in the bulk chalcopyrite
XPS indicated that polysuldes did not play a role in inhibiting (the (001)-S, (100)-S, and (112) surfaces), the formation of disulde
chalcopyrite dissolution, leading to the conclusion that chalcopyrite groups via an oxidative process with concomitant reduction of Fe(III)
dissolution occurred via the oxidation of the disulde phase (Parker to Fe(II) or, for the sulfur-terminated (111)-S face, the formation of the
et al., 2003). In a subsequent detailed study of a freshly-cleaved S2
4 group; (ii) on surfaces terminated in metal atoms (the (001)-M,
chalcopyrite surface prepared in an inert atmosphere, Klauber (2003) (100)-M and (111)-M surfaces), the formation of metalmetal bonds
again identied a disulde phase and presented a well-constructed resulting in alloy-like structures beneath a layer of exposed sulfur
argument for the formation of a two-layer surface phase containing atoms, and (iii) the relaxation of metal atoms with downward migration
pyritic FeS2 at 50% of the density of bulk pyrite. He proposed a model (the (101), (110) and part of the (112) surfaces). The experimental
that described the physical and redox mechanisms of the reconstruction detection of the disulde groups has been reported (e.g., Klauber,
and concluded that it would lead to a lower-energy chalcopyrite 2003) but the predicted formation of the proposed metal metal, alloy-
surface. like groups and/or metal migration requires greater investigation in
de Oliveira and Duarte (2010) used density functional theory within respect of elucidating the slow dissolution kinetics of chalcopyrite.
plane wave framework to analyse the reconstruction of the chalcopyrite de Lima et al. (2011) directed their study towards water adsorption
(001) surface and its reactivity, at the molecular level. The calculations on the reconstructed (001)-S and (001)-M chalcopyrite surfaces. Their
predicted that the sulfur-terminated (001) surface underwent a calculations predicted that (i) an iron atom was the most favourable
reconstruction in which disulde groups formed on the surface, site for a water molecule to adsorb on the reconstructed (001)-S surface,
consistent with the XPS results of Klauber (2003), with the concomitant (ii) dissociation of the water molecule was not favoured, and (iii) water
reduction of iron(III) to iron(II) oxidation state. adsorption on the (001)-M surface was also unfavourable, suggesting
Subsequently de Oliveira et al. (2012) applied plane wave density that the (001)-M surface would exhibit hydrophobic characteristics
functional calculations to a comparative study of the (001), (100), that might inuence surface reactivity. In an extension of that study,
(111), (112), (101) and (110) chalcopyrite surfaces including both the calculations of de Lima et al. (2012) revealed additional insights
sulfur- and metal-terminated cleavages of the rst three (Fig. 6). relevant to chalcopyrite leaching. Their calculations predicted that
Three different surface reconstructions emerged from the calculations (iv) water molecules and chloride ions, with their similar adsorption
of de Oliveira et al. (2012) which they described as follows: (i) on energies, competed for iron sites on (001)-S chalcopyrite surfaces,
surfaces which have relatively close sulfur atoms in the rst atomic (v) that sulfate and bisulfate ions were more strongly bound to the

Fig. 6. Chalcopyrite surface reconstructions which may partially impact on leaching kinetics.
Reprinted with permission from de Oliveira, C., de Lima G.F., de Abreu, H.A. and Duarte, H.A. (2012), Journal of Physical Chemistry C 116, 63576366. Copyright 2012 American Chemical
Society.
H.R. Watling / Hydrometallurgy 140 (2013) 163180 169

surface than water or chloride ion, that binding was bidentate, replacing during bioleaching. Third et al. (2002) and Gericke et al. (2010) also
two water molecules, and that bisulfate was bound more strongly than used oxygen limitation to control ORP during bioleaching, the former
sulfate, and (vi) that high acidity would lead to the protonation of the by temporarily arresting the air supply to the reactor when the ORP
disuldes on the surface, probably weakening the disulde bond. It is was greater than a designated set point, and the latter by adjusting
anticipated that future contributions from this research group should impeller speed to control oxygen mass transfer. Ahmadi et al. (2010)
advance our understanding of the slow kinetics of chalcopyrite leaching, exercised electrochemical control of ORP by applying a direct current
especially where predicted surface changes and interactions lend into the ore suspension with/without bacteria.
themselves to experimental validation. Viramontes-Gamboa et al. (2007) studied the oxidative dissolution
Recently, Li et al. (2013) conducted a comprehensive review of the of chalcopyrite in acidied ferric/ferrous sulfate media in the tem-
literature on the leaching of chalcopyrite, specically the nature and perature range of 2580 C and concluded that chalcopyrite displayed
roles of secondary product overlayers, the reconstruction of the bulk the activepassive behaviour of passivating metals. They showed that
sulde crystal when fractured or otherwise exposed as a surface and the predicted electrochemical passivation potentials (E-pp) obtained
the consequences of reconstruction in different leaching systems, and using electrochemical techniques were consistent with the results
the participation of redox reactions (both oxidation and reduction) from leaching experiments and reported that E-pp increased from
during leaching. While they reviewed many of the publications 680 mV (versus SHE) at 25 C to 755 mV at 80 C. In their systematic
discussed by Klauber (2008) and Debernardi and Carlesi (2012), Li study of the effect of acidity in the range of 2100 g H2SO4 L1, they
et al. (2013) adopted a broader approach, taking account of research also showed that E-pp were insensitive to acidity in the temperature
using oxidants other than ferric ion and a consideration of the above ranges of 2540 C and 6080 C but that between 40 and 60 C, an
modelling predictions. Li et al. (2013) concluded that three important acid-dependent transition of E-pp occurred from 680 to 755mV (versus
questions remained to be answered in order to explain the chalcopyrite SHE). Subsequently, Viramontes-Gamboa et al. (2010) collated data
leaching mechanism(s), relating to: (i) the interaction between from studies in which chalcopyrite leaching kinetics were investigated
chalcopyrite surfaces and oxidants; (ii) the relationships between under conditions of controlled ORP and found that chalcopyrite
oxidants and secondary surface reaction products and (iii) elucidation passivation was reported to occur at one of two potentials, either
of the key factor that determines whether a chalcopyrite surface will greater than about 680 mV or greater than about 750 mV (versus SHE)
become passivated. for samples from different locations and with different impurities.
They then demonstrated electrochemically that transitions from active
to passive and vice versa did not take place at the same potential; at
2.4. Redox control in ferric sulfate systems
temperatures of 50 C or higher and conditions of increasing potential,
chalcopyrite passivated at about 750755 mV (versus SHE) but if the
The acceleration of copper extraction from chalcopyrite in acidic
chalcopyrite was already passivated and the potential was decreased,
ferric/ferrous sulfate media at low ORP in the temperature range of
the chalcopyrite only became active at about 680 mV (versus SHE).
2050 C (e.g., Fig. 7), is well established. As a consequence, different
Viramontes-Gamboa et al. (2010) noted that this result offered a
strategies for ORP control were implemented in fundamental studies
possible explanation for the apparent discrepancies between published
and process developments. Bruynestein et al. (1986) established a
controlled-ORP leaching studies but did not elucidate the factors that
slurry potential of 540660 mV versus SHE via the initial addition of
determined whether chalcopyrite would be active or passive during
thiosulfate and cupric ions in a process that was specically designed
leaching in the potential range of 680755 mV (versus SHE).
to promote the extraction of copper from chalcopyrite and the partial
The benets of enhanced chalcopyrite dissolution under conditions
oxidation of the sulde to elemental sulfur (not sulfate). Kametani
of controlled ORP were sufciently important to have prompted patent
and Aoki (1985) maintained a bioleaching suspension at the relatively
applications relating to stirred tank and heap technologies. The patent
low ORP of 635 mV (versus SHE) by controlled additions of potassium
of Bruynestein et al. (1986) was noted above. In the patent of Dixon
permanganate solution, to maximise copper extraction from a
and Tshilombo (2005), the chemical leaching of chalcopyrite in acidic
chalcopyrite concentrate. Sandstrm et al. (2005) maintained an ORP
sulfate medium, with added pyrite as catalyst, was conducted under
of 635 mV (versus SHE) during the chemical leaching of a chalcopyrite
conditions whereby the pyrite is not materially oxidised, for example
concentrate using potassium permanganate, but reduced the airow
by maintaining the operating solution potential at a suitable level.
(oxygen) and added sodium sulte to maintain a similarly low ORP
Redox-control during chalcopyrite bioleaching was also a key claim in
patents by van der Merwe et al. (1998), Pinches et al. (2001a, b) and
100 Lastra and Budden (2002). Pinches et al. (2001a) proposed to bioleach
bacterial ORP control chalcopyrite in stirred tanks by controlling air, oxygen and/or carbon
dioxide concentrations to limit the efciency of iron(II)-oxidation by
80 the microbial community, specically to control the ORP in the range
copper extraction (%)

bacterial 575675 mV (versus SHE) thereby to promote the rate and extent of
chalcopyrite oxidation. Pinches et al. (2001b) claimed that the above
60
methodology could also be applied in a heap leaching process in
which the leach solution was conditioned to provide the said surface
40 potential. This latter claim was supported by the results of comparative
column leaching tests in which the iron(III)H2SO4 feed solutions were
chemical ORP control
produced in a separate bacterial ferric-ion-generator and could thus be
20 controlled at high redox (775 mV versus SHE) or low redox
(637 mV), as required. The authors noted that in a normal heap leach
chemical system, the presence of the natural bacterial community would make
0
control of ORP extremely difcult. Lastra and Budden (2002) proposed
0 5 10 15 20 25
two methods of controlling the ORP in heap bioleaching of chalcopyrite.
leach duration (days) In their invention, they described (i) the direct method: the im-
Fig. 7. Enhanced copper extraction at 50 C from chalcopyrite with bacteria and redox
provement of maintaining the reduction potential at below or around
control (re-drawn from Ahmadi et al., 2010). Electrochemical control of ORP at EH b 550 mV (reference electrode not specied but probably SCE or
400450 mV versus Ag/AgCl reference electrode. Ag/AgCl), a condition in which it was proposed could be achieved by
170 H.R. Watling / Hydrometallurgy 140 (2013) 163180

limiting the concentrations of oxygen in a heap operated at N 50 C; and oxidation (e.g., Mazuelos et al., 1999; Okibe and Johnson, 2002; Torma
(ii) the indirect method, in which they proposed that biological and Itzkovitch, 1976).
oxidation of ferrous ions would be achieved in a separate reactor and
0
that oxygen would be excluded from the heap, so as to prevent ferrous 2S 3O2 2H2 O 2H2 SO4 : 6
ion oxidation in situ. Key claims in the patent of Hunter (2006) were
that a chalcopyrite heap would be inoculated with a sulde-oxidising 2.5.2. Bioleaching at moderate to high temperature
bacterial culture that was either inefcient at oxidising iron(II) or did Given the strong temperature dependence of chalcopyrite dis-
not have that specic ability, and that the process water being fed to solution (Fig. 8), the bioleaching of chalcopyrite at temperatures
the heap would be maintained at ORP lower than 695 mV (versus 6090 C is an attractive R&D goal (e.g., Crdoba et al., 2008d; Gericke
SHE), such that the prevailing chemical conditions are conducive et al., 2001; Rodrguez et al., 2003) but not yet a full-scale, commercial
to leaching chalcopyrite while being non-conducive to surface process (e.g., Batty and Rorke, 2006). The iron(II)-(reaction (3)) and
passivation. As heaps are open systems, the maintenance of low-ORP sulfur-oxidising (reaction (6)) capabilities of archaeal members of
would be difcult should adventitious colonisation of the ore by native Acidianus, Metallosphaera and Sulfolobus and possibly other genera are
iron(II)-oxidising microorganisms occur. exploited in higher-temperature bioleaching. Some recent key ndings,
most often arising from studies using individual species but applicable
2.5. Microorganisms as catalysts in ferric sulfate systems to other archaea, are summarised: Archaeal species contribute to stable
leach environments in continuous pilot plants operated at 1220%
Bioleaching can be considered as a chemical system catalysed by solids loadings (Gericke and Pinches, 1999; Sandstrm and Petersson,
microorganisms. The bioleaching of sulde minerals with emphasis on 1997). Limitations resulting from reduced gas solubilities in tanks
copper sulde leaching using low-temperature heap technology was operated at high temperatures (6080 C) can be minimised by
reviewed (Watling, 2006 and references therein). Briey, sulde supplying oxygen and/or carbon dioxide to the reactors (Batty and
minerals in pristine (but exposed) deposits, mine waste dumps Rorke, 2006; Dew et al., 2001). Microbial sensitivity to shear from
and tailings, and/or managed heaps represent complex habitats for agitators results in an upper limit for solids loading of about 12.5% at
acidophilic microorganisms but very few of the microorganisms have demonstration scale (Batty and Rorke, 2006). Archaea exhibit different
been cultured, isolated or identied. sensitivities to soluble metals arising from the leaching of impure
concentrates or complex ores, but in the case of copper, archaeal
2.5.1. Ambient- to moderate-temperature bioleaching cultures tolerate up to 35 g Cu L1 (Batty and Rorke, 2006). While
Microorganisms participate in the dissolution of sulde minerals in redox-controlled bioleaching of chalcopyrite at temperatures 2050 C
two main ways: (i) by catalysing the oxidation of ferrous ions to ferric results in enhanced copper extraction, redox-control is less benecial
ions (reaction (3)) more rapidly than occurs in chemical systems and in the temperature range of 6080C (Gericke et al., 2010). The addition
(ii) by oxidising sulfur to sulfate, helping to remove sulfur overlayers of iron(II) (0.5 g L1) promotes cell growth, high ORP and increased
and generating additional acid (reaction (6)). The costs associated chalcopyrite dissolution rates but the presence of iron at N5 g L1
with the maintenance of sulde bioleaching microorganisms are promotes jarosite formation on chalcopyrite surfaces that hinders
minimal because they gain energy from the redox reactions (3) and further dissolution (Crdoba et al., 2008d; Sandstrm et al., 2005).
(6) (oxygen is usually supplied by air), acquire carbon for growth Thus operating conditions should be optimised to minimise the
from the carbon dioxide in air and obtain phosphorus, nitrogen, formation of iron-rich overlayers, possibly requiring strategies for iron
potassium and micronutrients from the ore environment. Microbial control. Cell attachment to surfaces is rapid (within 20 min) (Konishi
attachment to mineral surfaces is mediated by the properties of the et al., 1999, 2001). As long as a fraction of the population can attach
extracellular polymeric substances produced by the cells and to the chalcopyrite surface, copper extraction is efcient, but the
attachment is mineral and site specic and may change the properties population may not grow if cells are prevented from approaching the
of the underlying mineral surface (Watling, 2006 and references chalcopyrite surface (Gautier et al., 2008). Attached cells oxidise
therein). In some cases the addition of organic compounds may enhance elemental sulfur on particle surfaces (Liang et al., 2012) and planktonic
copper extraction (e.g., Hiroyoshi et al., 1995; Li and Li, 2010; Mallick cells oxidise the soluble reduced sulfur or iron(II) species that accu-
and Dasgupta, 1997) but, more generally, the presence of organic mulate in solution (Jordan et al., 2006). The joint activity of attached
compounds is deleterious to microbial growth, iron(II)- and/or sulfur cells and planktonic cells has been termed cooperative leaching

120
archaea

100
copper extracted (%)

80
in 96 hours

60
moderate
thermophiles
40
mesophiles

20

0
25 30 35 45 50 55 60 65 70 75 82
temperature (degrees Celsius)

inoculated tests abiotic tests


Fig. 8. Temperature dependence of chalcopyrite dissolution. Copper extraction from a chalcopyrite concentrate of composition 64% chalcopyrite, 6.6% pyrite, 3.3% pyrrhotite and 25%
quartz, with mean particle size 27 m, P80 66 m and surface area 0.36 m2 g1 (5-point BET analysis). Tests (96 h) inoculated with single species or mixed cultures; 2535 C, mesophiles;
4555 C, moderate thermophiles; and 6082 C, thermophiles (archaea). Abiotic tests were not inoculated.
H.R. Watling / Hydrometallurgy 140 (2013) 163180 171

(Rodrguez et al., 2003). Archaeal activity does not change the others in which sulfur addition was subsequently dened to mean
underlying mechanism of chalcopyrite dissolution by ferric ions. the addition of mineral suldes. This patent differed also in that the
However, surface-attached archaea may increase the corrosion primary purpose of the addition was to achieve a sufcient sulfur
potential and corrosion current density of chalcopyrite and may also content (between 2 and 20% w/w) to promote a rapid increase in the
reduce the polarisation resistance of chalcopyrite during bioleaching internal temperature of a chalcopyrite heap, such that thermophilic
(Li and Huang, 2011). microorganisms could grow and function. The concomitant production
of acid in the heap from microbial sulfur oxidation was noted as
2.5.3. Reduced sulfur additives to increase extraction benecial.
The addition of elemental sulfur to sulde bioleach operations is a In a laboratory study, augmentation of a chalcopyrite concentrate
strategy that can be used to meet three needs: (i) an alternative with elemental sulfur was intended to selectively promote the growth
means of providing the acid required to extract the target metals of sulfur-oxidising microorganisms in an iron- and sulfur-oxidising,
(reaction (6)); (ii) the generation of heat to increase copper extraction mixed-mesophilic culture during bioleaching, the rationale being that
rates; and (iii) a means of promoting microbial colonisation in a low- sulfur overlayers formed on chalcopyrite surfaces could become rate
grade sulde ore. limiting if not removed efciently. However, Xia et al. (2012) found
In the usual method of heap leaching, sulde oxidation occurs in that the strategy conferred an advantage on Acidithiobacillus thiooxidans
acidic ferric sulfate solutions, so the presence of acid is essential. at the expense of Leptospirillum ferriphilum to such an extent that
However, for heaps of low-grade ore, greater than 99% of the bulk ore iron-oxidation was inhibited and copper extraction was reduced.
is comprised of gangue minerals which also consume acid, to a greater Optimisation of the conditions should correct the imbalance between
or lesser extent, with time. This acid consumption during gangue sulfur- and iron-oxidation. This is essentially the underlying concept
mineral dissolution constitutes a reagent loss because it does not of the Bioheap process (Hunter, 2002a,b) in which a sulfur-oxidising
contribute to the target-metal extraction. For the majority of ores culture inefcient at iron oxidation was used to inoculate a heap of
processed using heap technology, acid consumption is a major chalcopyrite-containing ore. The process also provided for iron control
operating cost. It is not surprising, therefore, that the biological by passage of process solution through a second heap of largely barren
production of sulfuric acid through the oxidation of suldes or and inert ore, thus reducing the possible formation of passivating
elemental sulfur has been described. Young et al. (2003) patented a iron(III) oxides/hydroxides/ hydroxysulfates on chalcopyrite surfaces.
method for microbiological production of acid in which two Uhrie et al. (2012) proposed the use of elemental sulfur to promote
applications (processes) were described, one having many of the microbiological growth in a separate bioreactor for the purpose of
characteristics of a heap leach of sulde minerals and the other generating an inoculum for a heap leaching process.
employing reactor technology, such as an agitated tank. In both, the
aim was to oxidise elemental sulfur or mineral suldes to produce 2.5.4. Separation of biological and chemical processes
sulfuric acid using sulfur-oxidising acidophilic microorganisms that The benets of implementing an indirect bioleaching process are:
tolerate the high sulfate and metal concentrations occurring in process (i) the independent optimisation of the biological and chemical
solutions. In the patent described by Duyvesteyn et al. (2002), a primary processes, (ii) restricted access of microorganisms to sulfur reaction
claim was that the process described would lessen or eliminate the need products, minimising sulfur oxidation to sulfate (reaction (6)) and, in
to supply commercially-purchased or chemically-produced sulfuric some reactor congurations, (iii) lower oxygen requirement during
acid (such as from an on-site acid production plant). The proposed leaching, (iv) a degree of acid and iron control through iron hydrolysis
technology/application was broad; the sulfur-amended ore could be in and jarosite formation (reaction ((5)) and (v) lower exposure of
the form of a slurry, a heap, or a charge in a vat, with or without prior microorganisms to potentially toxic components in concentrates or
agglomeration. Bouffard et al. (2009) listed the benets of sulfur ores.
addition directly to a bioleaching heap as being reduced acid con- The separation of biological and chemical processes during
sumption by gangue minerals, faster mineral leaching kinetics, and bioleaching was briey mentioned above in the context of ORP control
heat generation. However, they also noted the disadvantages of adding and is generally referred to in the literature as indirect bioleaching. It
commercially-available sulfur, which is sold and transported in prill, depends upon the propensity of acidophilic microorganisms to become
slate or pellet form. In these commercial products, particle sizes attached to inert substrates and form biolms, within which bacterial
range from 1 to 13 mm and, as such, have low surface areas. Sulfur is oxidation of ferrous ion occurs (reaction (3)) and from which the
virtually insoluble in water, the solubility of rhombic sulfur being circulating ferric-ion rich medium is fed to a chemical leaching reactor.
1.9 0.6 108 mol S8 kg1 (Boulegue, 1978). Therefore the thrust of Ferrous ion oxidation can be achieved with different types of reactors
the Bouffard et al. (2009) study was the addition of lignosulfonate to furnished with different inert support materials for the biolm. Most
improve the wettability of sulfur particles during a wet-grinding applications apply packed beds, uidised beds or rotating biological
process. In laboratory tests, Salo-Zieman et al. (2006) studied the contactors furnished with a variety of support materials (e.g., resin,
bioleaching of a pyrrhotite-rich, acid-consuming nickelcopper ore glass or alginate beads, activated carbon, ground silica, jarosite, poly-
and reported that the addition of elemental sulfur reduced the need ethylene, polyvinyl chloride, polystyrene). Relatively small (economic)
for acid supplementation during bioleaching with mesophiles. reactors can be used because ferric ion productivity is generally high. A
Kohr et al. (2004) described a process for heating a heap rapidly to number of two-stage processes have been developed and implemented,
temperatures at least 50 C by augmenting the fuel content of the ore most providing suitable growth conditions for At. ferrooxidans, the species
to at least 10 kg of exposed sulde per tonne of ore, via the addition of thought to be most prevalent in bioleaching at the time of those studies.
sulde minerals. They further dened the process as the addition Acronyms used to describe three examples are BACFOX (bacterial lm
preferably of small particle sizes of sulfur-containing materials that oxidation Karavaiko, 1985), BFIG (bacterial ferric ion generator Van
would generate a large amount of heat when oxidised. The details of Staden, 1998) and IBES (indirect bioleaching with effects separation
the process were as already described in a family of patents on the Carranza et al., 1993).
GEOCOAT process (Harvey and Bath, 2007), in which host rock was
coated with ne particles of material rich in sulde, and leached in a 2.6. Cations as catalysts in ferric sulfate systems
heap. Essentially the same process was described in their more recent
patent (Kohr et al., 2011). Of the patents examined thus far, the process Several studies have been conducted in which selected cations were
described by Hunter and Williams (2006) was the only one in which the added to sulfuric acidferric sulfate leach systems and the enhancement
addition of elemental sulfur was specically proposed, compared with (or not) of copper extraction from chalcopyrite reported. Barriga
172 H.R. Watling / Hydrometallurgy 140 (2013) 163180

Mateos et al. (1987) compared a number of elements and showed that The benets of pre-treating pyrite with silver and then using the
the precipitation of silver sulde onto fresh chalcopyrite surfaces, or treated pyrite in the Galvanox process were described by Nazari
onto already-passivated surfaces, reactivated chalcopyrite. In other et al. (2012). In that process, in the absence of silver-treated pyrite,
comparative tests, silver ions were the most efcient at promoting the surface layer of sulfur that forms on chalcopyrite particles is non-
chalcopyrite dissolution (Ballester et al., 1990, 1992; Escudero et al., conductive and limits electrical contact between the pyrite and
1993; Muoz et al., 2007a). Johnson et al. (2008) reported that chalcopyrite particles. However copper extraction in the Galvanox
concentrates containing high silver contents were amenable to process is enhanced when silver-treated pyrite is used. Nazari et al.
bioleaching at low temperatures with mesophiles but that concentrates (2012) proposed that the mechanism of silver-enhanced extraction
with lower silver contents required higher temperatures and moderate involved a small part of the added silver reacting with the sulfur layer
thermophiles. to increase its conductivity and facilitate the transfer of electrons
The benecial action of silver (Ag+) was thought to proceed through between chalcopyrite and pyrite particles. The pyrite (plus added silver)
reaction with the chalcopyrite surface with the formation of silver is recycled in the Galvanox process.
sulde (Ag2S) on the sulde mineral surface (reaction (7)) followed Bismuth has received less attention than silver as a copper leaching
by the oxidation of Ag2S by ferric ions (reaction (8)) (Miller and catalyst. Hiroyoshi et al. (2007) ranked bismuth second, and Escudero
Portillo, 1979). However, this simple mechanism does not account for et al. (1993) ranked bismuth fourth most-effective catalyst of the
all the observations made then and subsequently. For example, it was metal ions tested. In contrast, Muoz et al. (2007a) reported that
shown that ferric ion oxidation of Ag2S (reaction (8)) required bismuth did not enhance the leaching of copper from a sulde ore and
temperatures greater than 100 C (Dutrizac, 1994) and elemental sulfur Gmez et al. (1999) reported only a small positive effect. Mier et al.
was not detected on a chalcopyrite surface post silver-catalysed (1994) suggested that the mechanism involved the reaction of
leaching (reactions (8)and (9)) (Parker et al., 2003). Hypotheses bismuth with the phosphate added as part of the bioleaching medium
about the formation of a small amount of silver (Muoz et al., 1998; and precipitation of bismuth phosphate, preventing ferric ions from
Price and Warren, 1986) included reactions (9)((11)] (Kolodziej and complexing with the phosphate and thus increasing the oxidising
Adamski, 1984; Price and Warren, 1986). In addition, the formation of potential of the Fe(III)/Fe(II) couple.
Ag2SO4 in the presence of a large excess of Ag+ (Warren et al., 1984)
was attributed to high oxidation potentials and excess Ag+ near the 3. Sulfuric acid alternative oxidants
Ag2S surface (Price and Warren, 1986), a reaction that would deplete
Ag+ concentrations and result in lower (catalysed) copper extraction, The sulfuric acidferric sulfate system for the oxidative leaching of
as was also reported by Crdoba et al. (2008c, 2009). In the reaction chalcopyrite has been studied widely, and the redox chemistry, reaction
mechanism proposed by Hiroyoshi et al., 2002 for systems operated at products, and consequences of adopting certain controlled conditions
low solution potential, chalcopyrite was reduced to chalcocite and silver are well understood. The reagents are inexpensive and, with biological
ions reacted with hydrogen sulde (overall reaction (12)) and assistance, can be regenerated during processing. However, the
chalcocite was oxidised by ferric ions more readily than chalcopyrite oxidising potential of ferric ion is not particularly high and several
(reaction (13)). Hiroyoshi et al. (2007) reported that Ag+ caused the stronger oxidants have been tested for their ability to extract copper
critical potential (that potential below which chalcopyrite leaching from chalcopyrite. Half reactions with standard reduction potentials
occurs faster) to rise. are summarised for selected oxidants (Table 3) and copper extraction
data compared in Fig. 9. It should be noted that the cupric/cuprous
2 2
CuFeS2 4Ag 2Ag2 Son chalcopyrite surface Cu Fe 7 couple does not play a role in sulfate systems as concentrated as those
used in oxidative hydrometallurgy, as sulfate anions do not complex
3 0 2
and thus stabilise the cuprous ion, which is rapidly oxidised in
2Fe Ag2 S 2Ag regenerated catalyst S 2Fe 8 oxygenated or ferric ion systems.

2 2 0 0 3.1. Sulfuric aciddichromate


CuFeS2 4Ag Cu Fe 4Ag 2S 9

The dissolution of a chalcopyrite concentrate in medium containing


2 0 3
Ag Fe Ag Fe 10 0.2 M dichromate and 0.5 M sulfuric acid showed strong temperature
dependence in the range of 3080 C (Antonijevic et al., 1994). In
0
2AgCl Ag Cl2 photochemical decomposition 11 those tests, the consumption of dichromate was greater than would
be expected if elemental sulfur was the sole product of sulde oxidation
2 0 (reaction (14)) indicating that some of the sulde was oxidised to
2CuFeS2 4Ag Cu2 S 2Fe 2Ag2 S S 12

3 2 2 0
Cu2 S 4Fe 2Cu 4Fe S : 13
Table 3
In tests using ore particles in columns, simulating heap leaching, the Selected oxidants, half reactions and standard reduction potentials.
benet of silver catalysis was less clear. Silver addition resulted in a Handbook of Chemistry and Physics, 89th edition, CRC Press, 2008.

sharp but transient increased copper extraction rate from chalcopyrite Oxidant Half reaction Standard reduction
contained within in a complex ore (Ahonen and Tuovinen, 1990) but potential
resulted in improved copper recoveries from low-grade chalcopyrite (E0) [V]

ore in 300-day tests (Muoz et al., 2007b). In those process develop- Cupric ion Cu2+ + e Cu+ 0.153
ments where silver catalysis was shown to be effective, there remained Ferric ion Fe3+ + e Fe2+ 0.771

Dichromate Cr2O2 +
7 + 14H + 6e 2Cr
3+
+ 7H2O 1.232
a concern about the economics of adding silver to full scale operations.
Chlorate ClO +
3 + 6H + 6e 3H2O + Cl

1.451
Hu et al. (2002) suggested the use of silver-bearing concentrates as MnO + 2+
Permanganate 4 + 8H + 5e Mn + 4H2O 1.507
a cost-effective strategy. In the IBES process (Carranza et al., 1997; Hydrogen peroxide H2O2 + 2H+ + 2e 2H2O 1.776
Palencia et al., 1998) and the BRISA process (Romero et al., 2003), the Ozone O3 + 2H+ + 2e O2 + H2O 2.076

ow sheets included a methodology to recover and recycle silver from Peroxodisulfate S2O2 +
8 + 2H + 2e 2HSO4 2.123

S2O2
8 + initiator 2SO4 2.6
the leached residues.
H.R. Watling / Hydrometallurgy 140 (2013) 163180 173

90 0.1 M Fe(III) in H2SO4; 80C


80 0.1 M Fe(III); 0.18 M H2SO4; 90C
70 1 M Fe(III); 0.25 M H2SO4; 90C

Cu (% extracted)
60 0.2 M Cr2O7; 0.5 M H2SO4; 70C
50 0.1 M Cr2O7; 0.4 M H2SO4; 70C

40 0.5 M ClO3; 1 M HCl; 65C

30 20% H2O2; 0.1 M H2SO4; 70C


2 M H2O2; 3 M H2SO4; 40C
20
3 M H2O2; 1 M H2SO4; 25C
10
2.5% O3; 0.5 M H2SO4; 22C
0
0 10 20 30 40 50 60 0.2 M S2O8; 0.01 M Fe(III); 0.1 M H2SO4; 23C
time (hours)

Fig. 9. Example copper extraction data for chalcopyrite concentrates of similar particle sizes: ferric sulfate systems (Ferreira and Burkin, 1976; Jones and Peters, 1976; Munoz et al., 1979),
dichromate (Antonijevic et al., 1994; Aydogan et al., 2006), sodium chlorate (Xian et al., 2012), hydrogen peroxide (Adebayo et al., 2003; Antonijevic et al., 2004; Olubambi et al., 2006),
ozone (Havlik and Skrobian, 1990); peroxodisulfate (Dakubo et al., 2012).

sulfate (reaction (15)). The results were consistent with those of earlier 3.3. Sulfuric acidpermanganate
studies (Murr and Hiskey, 1981; Shantz and Morris, 1974). The
concentration of dichromate did not affect the dissolution rate and it Permanganate ion has been used to maintain solution ORP in the
was concluded that the chalcopyrite surface was covered by Cr(VI) acid leaching of chalcopyrite and pyrite but not as the primary oxidant.
(Antonijevic et al., 1994; Murr and Hiskey, 1981). Dichromate activity Kametani and Aoki (1985) and Sandstrm et al. (2005) used potassium
increased with increasing acid concentration in the range 0.252.0 M permanganate to control ORP during ferric sulfate chemical leaching of
H2SO4 but there was minimal additional benet when acid con- nely-ground chalcopyrite concentrates at 90 and 65 C, respectively.
centrations N 2M were used. The presence of chloride ions in the system From their results, Sandstrm et al. (2005) concluded that the
decreased copper recovery, an effect that was attributed to the permanganate oxidised ferrous ions to ferric ions (reaction (17)), rather
displacement of Cr(VI) ions with Cl ions on the chalcopyrite surface. than oxidising the chalcopyrite directly. They noted that a disadvantage
Concentrations of ferric ions (up to 0.09 M) and copper ions (up to of using potassium permanganate was the loss of ferric ions from
0.08 M) had negligible effects on reaction rates. solution via reaction with potassium and sulfate ions to form jarosite
(reaction ((5)) resulting in lower copper extraction. Potassium perman-
6CuFeS2 5K2 Cr2 O7 35H2 SO4 6CuSO4 3Fe2 SO4 3 5K2 SO4 ganate was also a suitable oxidant for controlling ORP during the ferric
0
12S 5Cr2 SO4 3 35H2 O 14 sulfate leaching of pyrite, and the mass of potassium permanganate
solution required to maintain an ORP set point was a reliable indicator
6CuFeS2 17K2 Cr2 O7 71H2 SO4 6CuSO4 3Fe2 SO4 3 17K2 SO4 of leaching progress (Bouffard et al., 2006).

17Cr2 SO4 3 71H2 O: 15


2 3 2
5Fe MnO4 8H 5Fe Mn 4H2 O: 17
In the study by Aydogan et al. (2006), dissolution of a chalcopyrite
concentrate in H2SO4 (0.10.5 M) and K2Cr2O7 (0.010.15 M) media 3.4. Sulfuric acidhydrogen peroxidehydroxyl radical
was temperature dependent, the greatest extraction being achieved at
97 C, the top temperature tested. At 70 C, about 70% copper was Hydrogen peroxide is a strong oxidising agent in acidic medium
extracted in 150 min in medium containing 0.4 M H2SO4 and 0.1 M (Adebayo et al., 2003). It has been used as a leaching agent for uranium
dichromate. While dissolution with dichromate was faster than with ores (Eary and Cathles, 1983) and various studies on the use of peroxide
ferric ions, it was nevertheless controlled by diffusion through a porous with zinclead concentrate, pyrite and sphalerite have also been
sulfur layer (reaction (14)). conducted (e.g., Antonijevic et al., 1994; Olubambi et al., 2006). A
suggested oxidation mechanism in dilute aqueous solutions of
3.2. Sulfuric acidchlorate hydrogen peroxide involved dissociation to form the hydroxyl radical
(reaction (18), catalysed by ferrous ions) that oxidised the sulde
Chlorate ion is a strong oxidising agent (Table 3) that has been used moiety of minerals (reaction (19)) (Adebayo et al., 2003; Lin and
to enhance chalcopyrite leaching in both hydrochloric acid and sulfuric Luong, 2004).
acid systems. Kariuki et al. (2009) studied chalcopyrite oxidation by
mixing 2 g concentrate with up to 3 g sodium chlorate and 30 mL 2 3
H2 O2 Fe HO OH Fe 18
of 10 g L1 H2SO4 and heating in sealed Teon-lined vessels in the
temperature range of 45200 C. The results showed that chalcopyrite
oxidation (reaction ((16)) increased with increased temperature up to 2 0
2HO 2S s 2S s H2 O 0:5O2 : 19
approximately 165 C.

6CuFeS2 17NaClO3 3H2 SO4 3Fe2 SO4 3 6CuSO4 16 Both Adebayo et al. (2003) and Antonijevic et al. (2004) used
chalcopyrite concentrates in their studies, while Olubambi and
17NaCl 3H2 O:
Potgieter (2009) conducted an electrochemical study focused on the
Sodium chlorate concentrations up to 1 M were tested at 45 C with mechanism of chalcopyrite oxidation with acidic peroxide solution.
stirring in batch reactors (Xian et al., 2012). In these tests, the leaching Olubambi and Potgieter (2009) noted that chalcopyrite leaching with
rate accelerated with increased acid up to 1.5 M. At 0.5 M and 1 M HCl, sulfuric acid alone was slow (reaction (2)) but that, when hydrogen
respectively, 45% and 65% Cu were extracted in 300 min. peroxide was added to the system, copper extraction was rapid
174 H.R. Watling / Hydrometallurgy 140 (2013) 163180

and that the sulde was oxidised to elemental sulfur and sulfate solution and it has a high oxidising potential (Table 3). Havlik and
(reactions (20) and (21)). Skrobian (1990) studied the oxidation of a chalcopyrite concentrate
using an isothermal reactor containing 3 g sample and 1 L of 0.5 M
0
2CuFeS2 5H2 O2 5H2 SO4 2CuSO4 Fe2 SO4 3 4S 10H2 O H2SO4 into which ozone was introduced using a gas impeller. The
20 maximum ozone concentration was 3 vol.% in the gaseous stream
being sparged into the leachate. Copper extraction was efcient at
20 C but decreased at higher temperatures, an effect attributed to the
2CuFeS2 17H2 O2 H2 SO4 2CuSO4 Fe2 SO4 3 18H2 O: 21
lower ozone solubility. Because no elemental sulfur was detected in
any experiment and the ratio of Fe:Cu was 1:1, Havlik and Skrobian
Leaching rates were strongly dependent on temperature in a range
(1990) concluded that the reaction, comprising several possible
of 3080 C; with maximum copper extractions of 5% to 60% after
intermediate steps, proceeded according to overall reaction (23).
60 min. However, rates slowed after 6080 min at temperatures
Using results obtained at different temperatures and visualisation of
6080 C, attributed to increased peroxide at higher temperatures
leached surfaces, they proposed that the rate limiting step was the
(Adebayo et al., 2003). These authors also reported that best results
diffusion of ozone from the acidic medium to the chalcopyrite surface.
were obtained using 30% H2O2 v/v with H2SO4 and moderate agitation,
They also noted the possibility of using powerful ozonisers, which are
and suggested that high peroxide concentrations and strong agitation
used to prepare drinking water, in future mineral processing owsheets.
during leaching both caused accelerated hydrogen peroxide decom-
In subsequent research, Havlik et al. (1999) examined ways of making
position. Antonijevic et al. (2004) reported increased leaching rates in
the process more efcient, particularly in respect of the contact times
the range 0.15 M H2O2 in 2 M H2SO4 at 40 C. In both studies, higher
between gas bubbles and the mineral particles.
sulfuric acid concentrations resulted in higher dissolution rates (tests
conducted in the range 0.16 M H2SO4). More recently, Adebayo
(2006) showed that copper extraction from a low-grade chalcopyrite 3CuFeS2 8O3 3CuSO4 3FeSO4 : 23
ore was faster when leached in a solution (pH 4.3) containing ammo-
nium sulfate (72 g L1) with 45% v/v H2O2.
Carillo-Pedroza et al. (2010) studied the use of ozone as a strong
The addition of ethylene glycol to batch reactors containing nely
oxidant in an acidic ferric sulfate system applied to copper extraction
ground chalcopyrite in 1 M H2SO40.26 M H2O2 medium at 65 C
from an ore. They conducted isothermal agitated tests in which the
hindered peroxide decomposition, resulting in improved copper
solution was heated prior to ore addition (300 g in 1 L solution), and
extraction from 20% to 60% in 240 min (Mahajan et al., 2007). Similarly,
the oxygen/ozone mixture was injected into the base of the reactor
copper extraction of 65% was achieved in 60 min when ethylene glycol
(O3 0.5 or 1 g h1). The tests were carried out at 25 C, using a low-
was added to a 1.25M H2SO42.1M H2O2 solution but no further copper
grade, ground ore (0.7%Cu) split into different size fractions and leached
was extracted up to 300 min (Solis-Marcal and Lapidus, 2013).
in solutions of sulfuric acid (0.10.5 M) and ferric ions (00.5 M). Not
Olubambi et al. (2006) reported copper extraction of 74% in 180 min
surprisingly, best copper extractions were obtained using the smallest
from a ground, complex ore leached in 1 M H2O21 M H2SO4 solution
size fraction and highest ferric and acid concentrations, with 0.5 g h1
and attributed the increased dissolution rate with high peroxide
O3 inow. Overall, the results showed that ozone reduced the leaching
concentration to the formation of Caro's acid (reaction ((22)).
time and allowed the use of lower ferric ion and acid concentrations
when compared with a typical ferric sulfate leach system. It is also
H2 O2 H2 SO4 H2 SO5 H2 O: 22
worth noting that gangue minerals in the low grade ore apparently
did not consume ozone. However, the enhanced chalcopyrite leaching
Olubambi et al. (2006) suggested that optimisation of a system with
was less effective with larger particle sizes, from which result it was
1 M H2SO4 and peroxide concentrations up to 1 M could yield an
concluded that the presence of a single oxidant was sufcient to
economical process because the reagent costs were low and the overall
promote copper extraction from samples with low chalcopyrite surface
process would not be particularly corrosive in respect of materials of
areas. In their subsequent, broader study, Carillo-Pedroza et al. (2012)
construction. While the test was encouraging in that the peroxide was
summarised their results on the use of ozone as an alternative oxidant
active when an ore was used, this ore was atypical, comprising
for sulde ores as follows: Results show that the extraction of gold
massive chalcopyrite (66%) with minor pyrite (11%) and silica (7%). In
and silver is increased by at least 15%, with lower cyanide consumption;
an investigation of the kinetics of peroxide decomposition in the
the extraction of copper increased by 16% and in less time; and in the
presence of ground pyrite crystals in hydrochloric acid, Chiri (2007)
case of coal, sulfur is removed above 70%. They concluded that the
reported that peroxide decomposition was catalysed by both the iron
oxidation using ozone was a promising clean technology for processing
sites on pyrite surfaces and by aqueous ferric ions. Increases in pyrite
of sulde ores.
surface area, temperature, peroxide concentration and acid con-
centration all resulted in increased peroxide decomposition rates. The
highest losses of peroxide by decomposition into molecular oxygen 3.6. Sulfuric acidperoxodisulfate
and water were observed at moderate temperature (45 C) and initial
solution pH 2. In the presence of ferric ion ligands, only the mechanism The advantages of peroxodisulfate (S2O28 ) as an oxidising agent
of decomposition catalysed by the iron sites on the pyrite surface were reported to be that it could be made from sulfuric acid using an
contributed and thus overall decomposition was lower. The dissolution electrochemical cell (Canizares et al., 2005; Radimer and McCarthy,
of chalcopyrite involves the production of ferrous ions which would 1979) at low energy cost ($0.20 kg1) (Kimizuka et al., 2001).
be oxidised to ferric ions in leach solutions (reactions (20) and (21)). Peroxodisulfate undergoes decomposition or hydrolysis, depending
The additional ferric ions entering the solution would increase the upon the prevalent conditions (reactions (24) and (25)) but may also
decomposition rate of hydrogen peroxide (Garten, 1962; Kolthoff form reactive species and radicals (Table 3). In the leaching of chal-
et al., 1973), constituting a further reagent loss. copyrite using peroxodisulfate, the main activator for the formation of
sulfate radicals is expected to be the ferrous ion released from the
3.5. Sulfuric acidozone mineral during dissolution (Dakubo et al., 2012). In contaminated
aquifer remediation using ferrous ion-activated peroxodisulfate, the
Potential advantages of the use of ozone as an oxidising agent for decomposition rate of peroxodisulfate was 1020% per week (Brown,
chalcopyrite are that it does not introduce additional ions into the 2003). Once the peroxodisulfate has been converted to sulfate ions,
H.R. Watling / Hydrometallurgy 140 (2013) 163180 175

it can be regenerated electrochemically, minimising sulfuric acid copper extraction from chalcopyrite, reduced the amount of sulfate
consumption. formation and reduced the concentration of iron in solution. Similarly,
Deng et al. (2001) reported that the addition of small amounts of
2Na2 S2 O8 2H2 O O2 2H2 SO4 2Na2 SO4 24 chloride to a sulfuric acidO2 leach system enhanced copper extraction
from the copper residue produced from the oxidative leaching of nickel
Na2 S2 O8 H2 O NaHSO4 NaHSO5 : 25 matte.
In 1983 the Broken Hill Associated Smelters patented a process for
Dakubo et al. (2012) leached chalcopyrite concentrate or ore in the recovery of copper from a copperlead matte using oxygenated
stirred reactors or small columns at 23 C. When ground massive acidic sulfatechloride solutions (Sawyer and Shaw, 1983). According
chalcopyrite samples were leached in columns with solutions (pH 2) to Lu et al. (2000), this was possibly the only low-pressure sulde
containing sulfuric acid or peroxodisulfate (10 g L1) in sulfuric acid, leach process to have been commercialised. Lu et al. (2000) applied
copper extraction with peroxodisulfate was 500 times greater than the copper matte process to the dissolution of nely-ground chal-
without (% extraction not shown but roughly calculated to be copyrite concentrate in solutions of pH b 0.8 (0.8 M H2SO4) containing
approximately 10% extraction in 75 h). In tests using stirred batch 1 M NaCl in the temperature range of 6095 C. They achieved up to
reactors containing either 1.5 g of chalcopyrite concentrate or 20 g of 97% copper extraction in 9-hour tests. Based on their results showing
ground, low-grade ore in 1 L H2SO4 (pH 2), dissolved oxygen enhanced that chloride concentrations N0.5 M did not enhance the leaching rate
copper extraction but the pre-addition of ferrous ions to the reactors of chalcopyrite (consistent with the results of Palmer et al., 1981), Lu
had little effect. Up to 30% of the contained copper was extracted in et al. (2000) concluded that it was important only that there were
90 h when the ground ore was leached with 10 g L1 Na2S2O8 at pH 2. sufcient chloride ions present rather than an excess. This nding has
Up to 50% Cu was recovered from concentrate at 45 C using 50 g L1 commercial implications for geographical areas lacking in freshwater
Na2S2O8 at pH 2. Particle size inuenced extraction; extractions were but with a supply of brackish water or seawater; seawater contains
70% (25 m particle size) and 25% (125 m particle size), under similar approximately 0.5M chloride ions. From an examination of the residues,
leaching conditions. Dakubo et al. (2012) estimated a molar ratio of Lu et al. (2000) concluded that the sulfur reaction product obtained in
S2O2
8 added to Cu extracted as 8.8:1, for a 96-hour test with 50 g L1 the presence of 1 M NaCl was crystalline and porous, allowing reactants
Na2S2O8 (i.e., not reagent limited). to diffuse through the surface product layer to the unreacted mineral
surface. O'Brien et al. (1999) piloted a similar process involving oxygen
4. Hybrid sulfatechloride systems sparging of an agitated leach at 8095 C. Ground chalcopyritepyrite
ore (4% Cu) was used for the tests of duration less than 24 h. The pyrite
4.1. H2SO4NaClO2 in the ore promoted chalcopyrite oxidation via galvanic interaction and
up to 95% of the chalcopyrite was oxidised, compared with only 17% of
The advantage of using sulfuric acid and sodium chloride to create a the marcasite/pyrite content.
pseudo-chloride lixiviant is that these reagents are cheaper than ferric The Cuprochlor process (Espejo et al., 2001) has been used to extract
chloride or cupric chloride. The addition of chloride to a sulfate system copper from copper oxide and mixed copper oxide/secondary sulde ores
radically changes the solution speciation, as both iron and copper ions containing chalcocite, bornite and covellite (Herreros et al., 2006). In part,
can form complex species with chloride ions which may assist leaching. the process consisted of ne crushing of the ore and agglomeration with
The oxidant in this system is oxygen. There should be no need to add the addition of rst a calcium chloride solution (12 kg CaCl2 t1 ore) and
ferric ions to the system initially; ferrous ions produced during then sulfuric acid. This strategy resulted in the formation of competent
chalcopyrite leaching will be oxidised to ferric ions (reactions (2) and agglomerates bound by gypsum, as described by Vraar et al. (2000), as
(3)) and will subsequently contribute to overall copper extraction. well as high chloride concentrations in solution. After curing, the heaped
With careful selection of leaching conditions, most of the iron and ore was leached with high-chloride solution (90 g L1) containing
some sulfate can be precipitated as sodium jarosite (reaction ((5)). copper (5 g L1). At the Michilla mine in Chile (Antofagasta Minerals),
The addition of chloride to a sulfate system changes the speciation of 65% of the contained copper was extracted in 50 days of leaching and
both copper and iron through the formation of Cu- and Fe-complex 90% in 110 days. The process was considered to be competitive with
chloride ions in concentrations reecting the overall solution compo- bioleaching (Herreros et al., 2006). A modication of the Cuprochlor
sition. However, there is as yet no consensus on which complex species process to facilitate the leaching of copper sulde concentrates (including
are benecial to copper extraction under different conditions. Li et al. chalcopyrite) in heaps incorporated the calcium chloride agglomeration
(2010) examined the extraction rates of copper from a chalcopyrite strategy (Rauld Faine et al., 2005a), by which it was differentiated from
concentrate for a suite of leach systems and concluded that Fe3+ activity other processes for the leaching of concentrates in heaps.
was a key parameter. In tests conducted at 75 C, they measured faster In a separate development, Aroca Alfaro and Rauld Faine (2004)
kinetics for the NaCl (0.25 M)H2SO4 system at pH 2 (97% Cu extraction used a reactive chloridesulfate gel mixed with copper concentrate to
in 170 h) than at pH 1 (58% Cu extraction in 170 h). The slower rate at form a high viscosity paste which was poured into an open mould
pH 1 was attributed to the increased solubility of iron-containing that allowed the ingress of oxygen. The mixture was allowed to settle
secondary minerals due to the formation of FeCl2+ solution species and react, with the periodic addition of water to counter evaporation
and the consequent lower Fe3+ activities in the presence of chloride. and sulfuric acid to maintain sulfate activity. Copper was recovered
Ruiz et al. (2011) reported that leaching of a nely-ground chalcopyrite easily from the resulting dry paste, which contained the copper salts
concentrate in sulfatechloride solutions was rapid, 90% of the copper as chloride and sulfate compounds. A similar process to leach copper
being extracted in 180 min at 100 C. The presence of 0.5 M chloride from copper concentrates under pressure but at ambient temperature
ions (29 g L1 NaCl) enhanced the leaching rate signicantly but the using a reactive chloridesulfate gel was described by Rauld Faine
addition of 3 g L1 Fe3+ caused both an increase in the ORP and a et al. (2005b); the authors specically included chalcopyrite as one of
large decrease in the leaching rate. Muoz-Ribadeneira and Gomberg the target minerals.
(1970, 1971) attributed the increased copper extraction from
chalcopyrite when as little as 0.1 M HCl was added to a sulfuric acid 4.2. H2SO4Fe2(SO4)3NaCl.
leach to the in situ generation of the oxidant CuCl24 from the mixed
acid. Subramanian and Ferrajuolo (1976) leached a complex sulde The systems discussed in this section are similar to the above but are
ore in sulfuric acid with O2 over pressure (100 C) and reported that differentiated by the inclusion of ferric ions in leach media, rather
the presence of chloride ion (0.510gL1) resulted in slightly enhanced than reliance on their production during chalcopyrite oxidation. The
176 H.R. Watling / Hydrometallurgy 140 (2013) 163180

inuence of sodium chloride on ferric sulfate oxidation of chalcopyrite many years (Bardt, 19191922; Bjorling, 1968; Brennecke et al., 1981;
was studied at 95 C and atmospheric pressure using a nely-ground Davies et al., 1981; Habashi, 1973; Prater et al., 1973; Westby, 1918),
chalcopyrite concentrate (Carneiro and Leo, 2007). Solutions were but the only process to have been operated at commercial scale is
prepared by dissolving ferric sulfate and sodium chloride and adjusting the Nitrogen-Species-Catalysed (NSC) sulfuric acid pressure leach
the solution to pH0.15 with concentrated sulfuric acid. Using a 5% solids (Anderson, 1999, 2000). The advantage of nitric acid at moderate
loading, up to 90% of the copper was extracted with added sodium temperature is the absence of corrosion of stainless steel equipment.
chloride (58 g L1; 0.5 M) compared with 45% of the copper in the Disadvantages are that the industry has had relatively little experience
absence of sodium chloride (Carneiro and Leo, 2007). In that study, of the system and that nitrate in leach solutions may present problems
sodium chloride addition was reported to enhance chalcopyrite dissolu- for SX systems because the nitrogen species may degrade extractants.
tion by reducing the total iron and ferric iron concentrations via jarosite
precipitation (reaction ((5)), forming cuprous chloride complex ions 5.1. Sulfuric acidnitric acid
and enabling the participation of the Cu(I)/Cu(II) redox couple in
leaching, and increasing the surface area and porosity of the sulfur Queneau and Prater (1974) patented a process in which a sulde
reaction product. In another study (Kinnunen and Puhakka, 2004), a material (including copper sulde) was subjected to leaching in nitric
much lower concentration of sodium chloride (5 g L1) resulted in acid solution under conditions controlled to ensure the formation of
enhanced chalcopyrite (bio)leaching in ferric sulfate media; copper easily lterable hydrogen jarosite or other insoluble iron compounds.
extractions of about 100% and 80% were achieved at 87 C in 20 days Elemental sulfur was formed and reported with the residues and the
with and without added sodium chloride, respectively. However, the precipitated iron compounds in the lter cake. Unreacted sulde
enhancement in copper extraction with sodium chloride was much particles could be separated by froth otation after removal of the
less in similar 20-day tests at 68 C. In the tests with added sodium elemental sulfur and recycled to the leach system. The testwork
chloride, solutions were pH b1 compared with pHs 1.21.3 in the associated with this development was described by Prater et al.
absence of sodium chloride, a factor inuencing the soluble iron(III) (1973), who suggested that the dissolution proceeded according to
concentrations (higher with sodium chloride) and amounts of jarosite reaction (26). It was noted that acid concentration was an important
formed (higher without sodium chloride). The results of the above parameter, strongly inuencing the ratio of NO/NO2 generated from
recent studies were consistent with those of Dutrizac (1981) who nitric acid; the NO species is the more reactive and dominates solutions
reported that the addition of chloride ion to a ferric sulfatesulfuric b25% HNO3. Both temperature and nitric acid concentrations inuenced
acid system progressively accelerated the leaching of chalcopyrite and the amount of elemental sulfur formed (Prater et al. (1973)). The
promoted copper extraction signicantly. complexity of this system would necessitate careful testwork to
Dutrizac and MacDonald (1971) investigated the addition of sodium optimise copper extraction and elemental sulfur production.
chloride on the rate of chalcopyrite dissolution from low-grade (3% Cu)
copper sulde lump ore at low temperatures over prolonged periods, 6CuFeS2 22HNO3 9H2 SO4 6CuSO4 3Fe2 SO4 3 22NO 26
in a simulation of dump leaching conditions. In those tests, ore was 0
6S 20H2 O:
crushed to 6 mm and loaded into 25 mm-diameter columns to give a
1 m bed depth. The ore was leached with 0.1 M Fe3+ in 0.1 M H2SO4 Bjorling et al. (1976) aimed to develop a process for the treatment of
with/without 0.1 M NaCl without solution recycle. Extractions ranged chalcopyrite concentrate that met the criteria: (i) a closed circuit from
from 1 to 8% Cu in tests conducted in the temperature range of which products can be utilised or disposed in an environmentally
2540 C but, more importantly, the presence of sodium chloride friendly manner; (ii) air as the primary oxidant because of its
inhibited chalcopyrite dissolution. Dutrizac and MacDonald (1971) availability and low cost; and (iii) minimised construction costs by
concluded that, as dump leach temperatures are generally below 50 C, using simple equipment. The further condition, that the sulde moiety
large chloride additions would not be benecial. Nevertheless, in some of the chalcopyrite must be oxidised to elemental sulfur (not sulfate)
arid regions where freshwater supplies are restricted and water use in by controlling ORP required a secondary oxidant for the oxidation step
mineral processing must be efcient, seawater was substituted for that could be regenerated using air. They described the system as
freshwater in a few unit processes, including heap leaching of copper oxidation of chalcopyrite by oxygen (air) in sulfuric acid in the presence
ores. Clearly, the main benet to the mine was related to water efciency of a small amount of nitric acid, which acted as a catalyst (reaction (27)).
and there are few published data on the chemistry of the hybrid sulfate Nitric acid is an efcient oxidant, becoming reduced to nitric oxide
chloride system. Philippe et al. (2010) reviewed those operations using (reaction (28), in which the protons are supplied by the sulfuric acid);
seawater and noted that the majority had to overcome specic issues nitric oxide can be re-oxidised to nitric acid in the presence of oxygen
related to water quality, corrosion management and process efciency. and water (reaction (29)).

4.3. H2SO4Fe2(SO4)3LiCl 4CuFeS2 5O2 20H 4Cu


2 3
4Fe 10H2 O 8S
0
27

In tests using relatively dilute sulfate solutions, Dutrizac (1981) noted  


0
NO3 4H 3e NO 2H2 O E 0:957V 28
that the addition of lithium chloride in the concentration range of 14 M
converted the sulfate system (0.1 M Fe2(SO4)3; 0.3 M H2SO4; 90 C) to a
pseudo-chloride system. In that study, the addition of lithium chloride 4NO 3O2 2H2 O 4HNO3 : 29
was a means of increasing the chloride concentration without promoting
the formation of jarosite-like compounds, as happened when sodium or In the process developed by Bjorling et al. (1976), nitric acid and
potassium chlorides were added into sulfate systems. The addition of hydrogen ions had to be supplied continuously at the same rates as
chloride ion caused a gradual increase in the leaching rate, up to 2.6 they were consumed (pH 0.9) in order for the critical ORP to be
times at a concentration of 4 M chloride ion; however, greater chloride maintained. Under such conditions the formation of sulfate could be
concentrations resulted in the precipitation of lithium sulfate. minimised and other parameters optimised. The nitric oxide could be
re-converted to nitric acid outside the reaction vessel (reaction (29)).
5. Hybrid sulfatenitrate systems In this process, a retention time of 2.5 h resulted in almost complete
chalcopyrite dissolution at 95 C and, as in the Queneau and Prater
Nitric acid is both a strong acid and a strong oxidising agent. The (1974) process, the unreacted suldes were recovered by otation
potential use of nitric acid in copper processing has been studied over and recycled to the leach system.
H.R. Watling / Hydrometallurgy 140 (2013) 163180 177

0
Researchers at Kennecott Copper in Utah built a pilot plant for S 3NO2 H2 O H2 SO4 3NO 36
treating copper concentrate with a nitricsulfuric acid mixture and
concluded that the process was feasible but not economical because of NO 2NaNO3 H2 SO4 3NO2 Na2 SO4 H2 O 37
the formation of a small amount (0.5%) of nitrous oxide which cannot
be reconverted to nitric acid and thus represents a loss of reagent in 3NO2 H2 O 2HNO3 NO: 38
this process (Habashi, 1999). A further requirement with economic
consequences was that nitrous oxide must not be released to the
6. Summary
atmosphere unless rst reduced catalytically to nitrogen. Nitric acid is
expensive because it is produced from the oxidation of ammonia but if
A number of drivers exist for the development of different hydro-
it could be produced more cheaply from sodium nitrate, this might
metallurgical technologies for the extraction of copper from chalcopyrite.
inuence the economics in favour of a nitrate oxidative process for
They include the current imbalance between copper supply and demand,
chalcopyrite dissolution.
the overall decline in ore grades and the extensive exploitation of
low-grade oxide and secondary sulde ores that leave vast quantities of
low-grade chalcopyrite ores as a major, but thus far uneconomic, source
5.2. Sulfuric acidsodium nitrate
of copper. Other sources of copper are polymetallic ores of complex
mineralogy that cannot be concentrated and chalcopyrite concentrates
Sokic et al. (2009) examined the leaching of chalcopyrite in the
that contain impurities such as arsenic and cannot be smelted.
sulfuric acidsodium nitrate system. They chose leach conditions after
The primary focus of the review is on the chemistry of leaching in
consideration of the thermodynamics of the main chemical reactions
sulfate systems at atmospheric pressure. The use of chalcopyrite
in the temperature range of 2590 C. The matrix of tests involved as
concentrates for the majority of fundamental studies gives condence
variables: stirring rate, temperature, leaching time, particle size, and
that the chalcopyrite is entirely liberated and therefore exposed to
H2SO4 and NaNO3 concentrations. Solids loading in the tests was 20 g
the lixiviant/oxidant, provides a simple test matrix largely free of
concentrate in 1.2 L solution. Using conditions 80 C, 37 m particle
other phases that could mask the reaction chemistry, and accentuates
size fraction and 1.5 M H2SO4, copper extraction in 240 min was 40%
the impact of chalcopyrite grain size and surface area on leaching rates.
at 0.15M NaNO3; 60% at 0.3M NaNO3; and 75% at 0.9M NaNO3 addition.
There are many fundamental studies in which electrochemical
As Bredenhann and Van Vuuren (1996) had shown that sodium nitrate
and spectroscopic techniques have been used to probe chalcopyrite
oxidised millerite more rapidly than did ferric sulfate, Sokic et al. (2009)
surfaces, describe the leaching mechanism and identify the surface
did not make a similar comparison with ferric sulfate in their study of
species responsible for the passivation of chalcopyrite during leaching.
chalcopyrite dissolution. In a subsequent study, it was shown that
Attempts have been made in three recent substantial reviews to
elemental sulfur was formed during the reactions of chalcopyrite with
reconcile reported results, despite signicant differences between the
sodium nitrate (reactions (30)(33)) and that its presence on particle
conditions under which those results were obtained. One of the
surfaces slowed leaching (Sokic et al., 2010).
difculties of that reconciliation is highlighted by the recent prediction
of three different chalcopyrite surface structures that may leach via
3CuFeS2 4NaNO3 8H2 SO4 3CuSO4 3FeSO4 2Na2 SO4
0 different mechanisms at different rates. No reports on chalcopyrite
6S 4NO 8H2 O 30 leaching were found in which the surface orientation of the chalcopyrite
was identied prior to leaching, creating a future opportunity for focused
CuFeS2 4NaNO3 4H2 SO4 CuSO4 FeSO4 2Na2 SO4 research on chalcopyrite surfaces of known crystallographic orientation.
0
2S 4NO2 4H2 O 31 The range of solution chemistry parameters that inuence bulk
chalcopyrite dissolution in acidic ferric sulfate systems have been
studied comprehensively but often as independent variables. Many
6CuFeS2 10NaNO3 20H2 SO4 6CuSO4 3Fe2 SO4 3 5Na2 SO4
studies have generated results consistent with leaching rates being:
0
12S 10NO 20H2 O
32 largely independent of acid concentration above that required to
maintain a sufcient ferric ion concentration (which need not exceed
2CuFeS2 10NaNO3 10H2 SO4 2CuSO4 Fe2 SO4 3 5Na2 SO4 0.1 M Fe2(SO4)3),
0 dependent on sulfate concentration and solution ORP (with optimum
4S 10NO2 10H2 O:
ORP being dependent on ferricferrous and cupriccuprous ion
33 concentrations), and
The leaching of chalcocite in a sodium nitratesulfuric acid system enhanced by microorganisms (biological catalysts) or the addition of a
was studied by Vraar et al. (2003). For a system with conditions: 20 g chloride salt, under some conditions. Processes with the addition of
Cu2S in 1.2 L solution with stirring, temperature 80 C and duration of nitric acid or a nitrate salt are less well developed and the benets
leach 330 min and sulfuric acid 1 M, copper extraction increased less convincing.
with increased NaNO3 concentration from 30% at 0.15 M NaNO3 to
In studies using concentrates, chalcopyrite surfaces may become
95% at 0.6 M NaNO3. The system was as complex as the chalcopyrite
covered in overlayers of elemental sulfur or ferric oxy-hydroxy-sulfate
nitrate system in respect of chemical reactions. Both Vraar et al.
compounds such as jarosite. These bulk precipitates have often been
(2003) and Sokic et al. (2009) gave comprehensive accounts of the
blamed for the slower dissolution rates obtained after an initial period
reaction chemistry. The negative Gibbs free energy calculated for the
of relatively rapid copper extraction during leaching or bioleaching.
reactions (30)(34), (1), and (35)(37) showed that they were all
The effects may be overcome by changing the leaching mechanism
thermodynamically feasible at atmospheric pressure in the temperature
using a catalyst (e.g., Ag+) or by exploiting galvanic interaction with
range of 2590C. Only reaction (38) had a positive Gibbs Energy value.
another mineral (e.g., pyrite). Overlayers are unlikely to present a
6FeSO4 2NaNO3 4H2 SO4 3Fe2 SO4 3 Na2 SO4 2NO 4H2 O signicant problem in most whole-ore (heap) leaching systems unless
the amounts of iron(III) delivered through gangue dissolution are
34
excessive. Studies in which multiple variables are jointly optimised and
controls put in place to maintain optimum conditions, may be rewarded
0
S 2NaNO3 Na2 SO4 2NO 35 with increased copper extraction from chalcopyrite. While exercising
178 H.R. Watling / Hydrometallurgy 140 (2013) 163180

that degree of control in heaps is difcult but not impossible, other Bouffard, S.C., Tshilombo, A., West-Sells, P.G., 2009. Use of lignosulfate for elemental sulfur
biooxidation and copper leaching. Miner. Eng. 22, 100103.
reactor types should also be considered. In any style of reactor, a primary Boulegue, J., 1978. Solubility of elemental sulfur in water at 298 K. Phosphorus Sulfur 5,
requirement of whole ore leaching is the exposure of the chalcopyrite 127128.
grains to the selected lixiviant and oxidant. This may require special Bredenhann, R., Van Vuuren, C.P.J., 1996. The leaching behaviour of a nickel concentrate in
an oxidative sulphuric acid solution. Miner. Eng. 12, 687692.
ore preparation and/or, may exclude heap leaching as a process option. Brennecke, H.M., Bergmann, O., Ellefson, R.R., Davies, D.S., Lueders, R.E., Spitz, R.A., 1981.
The development of processes using sulfuric acid as lixiviant and Nitricsulfuric leach process for recovery of copper from concentrate. Min. Eng. 33,
oxidants other than ferric ions is an interesting research area that has 12591266.
Brown, R.A., 2003. In situ chemical oxidation: performance, practice and pitfalls. Paper
yet to be taken beyond laboratory scale. The main disadvantage is that Presented at the Air Force Center for Engineering and the Environment Technology
the alternative oxidants are more expensive than ferric ion. Countering Transfer Workshop (Source: http://clu-in.org/download/techfocus/chemox/4_brown.
the cost, the alternative oxidants are stronger oxidising agents pdf).
Bruynestein, A., Hackl, R.P., Lawrence, R.W., Vizsolyi, A.I., 1986. Biological acid leach
than ferric ions and copper extractions are much faster. Few studies
process. U.S. Patent 4,571,387 (18 February 1986).
have been undertaken using ores rather than concentrates. Thus the Canizares, P., Larrondo, F., Lobato, J., Rodrigo, M.A., Saez, C., 2005. Electrochemical
interactions of these reagents with gangue minerals have not been synthesis of peroxodisulfate using boron-doped diamond anodes. J. Electrochem.
described. Assuming one or more reagents can be found that are Soc. 152, D191D196.
Carillo-Pedroza, F.R., Snchez-Castillo, M.A., Soria-Aguilar, M.J., Martnez-Luvanos, A.,
selective for sulde minerals and sufciently aggressive to extract Gutirrez, E.C., 2010. Evaluation of acid leaching of low grade chalcopyrite using
copper from chalcopyrite, then a goal of future research may be to ozone by statistical analysis. Can. Metall. Q. 49, 219226.
develop a process in which a strong alternative oxidant can be Carillo-Pedroza, F.R., Soria-Aguilar, M.J., Pecina Trevino, T., Snchez-Castillo, M.A.,
Martnez-Luvanos, A., 2012. Treatment of sulde minerals by oxidative leaching
regenerated and recycled, thus minimising reagent costs. with ozone. Miner. Process. Extr. Metall. Rev. 33, 269279.
Carneiro, M.F.C., Leo, V.A., 2007. The role of sodium chloride on surface properties of
chalcopyrite leached by ferric sulphate. Hydrometallurgy 87, 7382.
Acknowledgements Carranza, F., Iglesias, N., Romero, R., Palencia, I., 1993. Kinetics improvement of high-grade
sulphides bioleaching by effects separation. FEMS Microbiol. Rev. 11, 129138.
Carranza, F., Palencia, I., Romero, R., 1997. Silver-catalyzed IBES process: application to a
The nancial support of the Australian Government through CSIRO
Spanish copperzinc sulphide concentrate. Hydrometallurgy 44, 2942.
Minerals Down Under Flagship is gratefully acknowledged. Chiri , P., 2007. A kinetic study of hydrogen peroxide decomposition in presence of
pyrite. Chem. Biochem. Eng. Q. 21, 257264.
Crdoba, E.M., Muoz, J.A., Blzquez, M.L., Gonzlez, F., Ballester, A., 2008a. Leaching of
References chalcopyrite with ferric ion. Part I: general aspects. Hydrometallurgy 93, 8187.
Crdoba, E.M., Muoz, J.A., Blzquez, M.L., Gonzlez, F., Ballester, A., 2008b. Leaching of
Adebayo, A.O., 2006. Dissolution of some sulphide minerals in the presence of hydrogen chalcopyrite with ferric ion. Part II: effect of redox potential. Hydrometallurgy 93, 8896.
peroxide solution and extraction of the base metals. PhD Thesis Federal University Crdoba, E.M., Muoz, J.A., Blzquez, M.L., Gonzlez, F., Ballester, A., 2008c. Leaching of
of Technology, Akure, Nigeria. chalcopyrite with ferric ion. Part III: effect of redox potential on the silver-catalyzed
Adebayo, A.O., Ipinmoroti, K.O., Ajayi, O.O., 2003. Dissolution kinetics of chalcopyrite with process. Hydrometallurgy 93, 97105.
hydrogen peroxide in sulphuric acid medium. Chem. Biochem. Eng. Q. 17, 213218. Crdoba, E.M., Muoz, J.A., Blzquez, M.L., Gonzlez, F., Ballester, A., 2008d. Leaching of
Ahmadi, A., Schafe, M., Mana, Z., Ranjbar, M., 2010. Electrochemical bioleaching of high chalcopyrite with ferric ion. Part IV: the role of redox potential in the presence of
grade chalcopyrite otation concentrates in a stirred bioreactor. Hydrometallurgy mesophilic and thermophilic bacteria. Hydrometallurgy 93, 106115.
104, 99105. Crdoba, E.M., Muoz, J.A., Blzquez, M.L., Gonzlez, F., Ballester, A., 2009. Comparative kinetic
Ahonen, L., Tuovinen, O.H., 1990. Silver catalysis of the bacterial leaching of chalcopyrite study of the silver-catalyzed chalcopyrite leaching. Int. J. Miner. Process. 92, 137143.
containing ore material in column reactors. Miner. Eng. 3, 437445. Dakubo, F., Baygents, J.C., Farrell, J., 2012. Peroxodisulfate assisted leaching of
Anderson, C.G., 1999. The treatment of chalcopyrite concentrates with nitrogen species chalcopyrite. Hydrometallurgy 121124, 6873.
catalysed oxidative pressure leaching. In: Young, S.K., Dreisinger, D.B., Hackl, R.P., Davies, D.S., Lueders, R.E., Spitz, R.A., Frankiewicz, T.C., 1981. Nitricsulfuric leach process
Dixon, D.G. (Eds.), Proceedings of Copper99Cobre99 International Conference. improvements. Min. Eng. 33, 12521259.
Hydrometallurgy of Copper, vol. IV. TMS, Warrendale, pp. 139149. de Lima, G.F., de Oliveira, C., de Abreu, H.A., Duarte, H.A., 2011. Water adsorption on the
Anderson, C.G., 2000. The industrial hydrometallurgical production of copper from reconstructed (001) chalcopyrite surface. J. Phys. Chem. 115, 1070910717.
concentrates with nitrogen species catalysed pressure leaching, solvent extraction de Lima, G.F., de Oliveira, C., de Abreu, H.A., Duarte, H.A., 2012. Sulfuric and hydrochloric
and electrowinning. ALTA SX/IX Forum. ALTA Metallurgical Services, Melbourne. acid adsorption on the reconstructed sulfur terminated (001) chalcopyrite surface.
Antonijevic, M.M., Jankovic, Z.D., Dimitrijevic, M.D., 1994. Investigation of the kinetics of Int. J. Quantum Chem. 112, 32163222.
chalcopyrite oxidation by potassium dichromate. Hydrometallurgy 35, 187201. de Oliveira, C., Duarte, H.A., 2010. Disulphide and metal sulphide formation on the
Antonijevic, M.M., Jankovic, Z.D., Dimitrijevic, M.D., 2004. Kinetics of chalcopyrite reconstructed (001) surface of chalcopyrite: a DFT study. Appl. Surf. Sci. 257, 13191324.
dissolution by hydrogen peroxide in sulphuric acid. Hydrometallurgy 73, 329334. de Oliveira, C., de Lima, G.F., de Abreu, H.A., Duarte, H.A., 2012. Reconstruction of the
Aroca Alfaro, F., Rauld Faine, J., 2004. Procedure to leach copper concentrates, under chalcopyrite surfaces a DFT study. J. Phys. Chem. C 116, 63576366.
pressure and at ambient temperature, by forming a reactive gel in a sulfatechloride Debernardi, G., Carlesi, C., 2012. Chemicalelectrochemical approaches to the study of
medium. European patent EP 1559799 (priority date 29 January 2004). passivation of chalcopyrite. Miner. Process. Extr. Metall. Rev. 34, 1041.
Aydogan, S., Ucar, G., Canbazoglu, M., 2006. Dissolution kinetics of chalcopyrite in acidic Deng, T., Lu, Y., Wen, Z., Liu, D., 2001. Oxygenated chloride-assisted leaching of copper
potassium dichromate solution. Hydrometallurgy 81, 4551. residue. Hydrometallurgy 62, 2330.
Ballester, A., Gonzlez, F., Blzquez, M.L., Mier, J.J., 1990. The inuence of various ions in Dew, D.W., Basson, P. and Miller, D.M., 2001. Recovery of copper from copper bearing
the bioleaching of metal sulphides. Hydrometallurgy 23, 221235. sulphide minerals by bioleaching with controlled oxygen feed. World Patent WO
Ballester, A., Gonzlez, F., Blzquez, M.L., Gmez, C., Mier, J.L., 1992. The use of catalytic 01/18269 (15 March 2001).
ions in bioleaching. Hydrometallurgy 29, 145160. Dixon, D.G., Tshilombo, A.F., 2005. Leaching process for copper concentrates. U.S. Patent
Bardt, H., 19191922. Recovering metals contained in metalliferous ore, waste residues Application 269208-A1 (8 December 2005).
and alloys. German Patent 353,795. Dixon, D.G., Mayne, D.D., Baxter, K.G., 2008. Galvanox a novel galvanically-assisted
Barriga Mateos, F., Palencia Perez, I., Carranza Mora, F., 1987. The passivation of atmospheric leaching technology for copper concentrates. Can. Metall. Q. 47, 327336.
chalcopyrite subjected to ferric sulfate leaching and its reactivation with metal Domic, E.M., 2007. A review of the development and current status of copper bioleaching
suldes. Hydrometallurgy 19, 159167. operations in Chile: 25 years of successful commercial implementation. In: Rawlings,
Basto, E., 2012. Escondida site tour. Source: http://www.bhpbilliton.com/home/investors/ D.E., Johnson, D.B. (Eds.), Biomining. Springer, Berlin, pp. 8195.
reports/Documents/2012/121001_Escondida%20Site%20Visit%20Presentation.pdf. Dopson, M., Lvgren, L., Bostrm, D., 2009. Silicate mineral dissolution in the presence of
Batty, J.D., Rorke, G.V., 2006. Development and commercial demonstration of the BioCOP acidophilic microorganisms: implications for heap bioleaching. Hydrometallurgy 96,
thermophile process. Hydrometallurgy 83, 8389. 288293.
Baxter, K., Dreisinger, D.B., Pratt, G., 2003. The Sepon copper project: Development of a Dreisinger, D., 2006. Copper leaching from primary suldes: options for biological and
owsheet. In: Young, C.A., Alfantazi, A.M., Anderson, C.G., Dreisinger, D.B., Harris, B., chemical extraction of copper. Hydrometallurgy 83, 1020.
James, A. (Eds.), Hydrometallurgy. TMS, Warrendale, pp. 14871502. Dutrizac, J.E., 1981. The dissolution of chalcopyrite in ferric sulfate and ferric chloride
Bjorling, G., 1968. Hydrometallurgical production of copper from activated chalcopyrite. media. Metall. Trans. B 12, 371378.
AIME Annual Meeting, New York, February 1968. Dutrizac, J.E., 1994. The leaching of silver sulde in ferric ion media. Hydrometallurgy 35,
Bjorling, G., Faldt, I., Lindgren, E., Toromanov, I., 1976. A nitric acid route in combination 275292.
with solvent extraction for hydrometallurgical treatment of chalcopyrite. In: Dutrizac, J.E., MacDonald, R.J.C., 1971. The effect of sodium chloride on the dissolution of
Yannopoulos, J.C., Agarwal, J.C. (Eds.), Extractive Metallurgy of Copper, vol. II. AIME, chalcopyrite under simulated dump leaching conditions. Metall. Trans. 2, 23102312.
NY, pp. 725737. Duyvesteyn, W.P.C., Budden, J.R. and Lastra, M.R., 2002. Recovery of metals from ore. U.S.
Bouffard, S.C., Rivera-Vasquez, B.F., Dixon, D.G., 2006. Leaching kinetics and stoichiometry Patent 6,387,239 (14 May 2002).
of pyrite oxidation from a pyritemarcasite concentrate in acid ferric sulfate media. Eary, L.E., Cathles, M., 1983. The kinetics of uranium dioxide dissolution in acidic
Hydrometallurgy 84, 225238. hydrogen peroxide solutions. Metall. Trans. B 14, 325334.
H.R. Watling / Hydrometallurgy 140 (2013) 163180 179

Escudero, M.E., Gonzlez, F., Blzquez, M.L., Ballester, A., Gmez, C., 1993. The catalytic Johnson, D.B., Okibe, N., Wakeman, K., Liu, Y., 2008. Effect of temperature on the
effect of some cations on the biological leaching of a Spanish complex sulphide. bioleaching of chalcopyrite concentrates containing different concentrations of silver.
Hydrometallurgy 34, 151169. Hydrometallurgy 94, 4247.
Espejo, R. Y., Arriagada F., D'amico, J., Jo, M., Rojas, J., Neuberg, H., Ruiz, M., Yanez, H., Jones, D.L., 1996. CESL copper process. ALTA Copper Hydrometallurgical Forum
Reyes, R., Bustos, S., Montealegre, R., Rauld, J., 2001. Procedimiento para aglomerar (Brisbane). ALTA Metallurgical Services, Melbourne.
minerales de cobre chancados no, mediante la adicion de cloruro de calcio y acido Jones, D.L. and Peters, E., 1976. The leaching of chalcopyrite with ferric sulfate and ferric
sulfurico y procedimiento de lixiviacion previa aglomeracion del mineral. Chilean chloride. In: Yannopoulos, J.C. and Agarwal, J.C., 1986 (Eds.), Extractive Metallurgy
Patent Application No. 40891 (19 May 1997; granted 2001). of Copper, Volume 2. AIME, New York, pp. 633 653.
Ferreira, R.C.H., Burkin, A.R., 1976. Acid leaching of chalcopyrite. In: Burkin, A.R. (Ed.), Jordan, H., Sanhueza, A., Gautier, V., Escobar, B., Vargas, T., 2006. Electrochemical study of
Leaching and reduction in hydrometallurgy. IMM, London, pp. 5466. the catalytic inuence of Sulfolobus metallicus in the bioleaching of chalcopyrite at
Fleury, F., Delgado, E., Collao, N., 2010. A new technology for processing hydro- 70 C. Hydrometallurgy 83, 5562.
metallurgical copper ore, Cobre Las Cruces Project. Proceedings of Copper 2010 Kametani, H., Aoki, A., 1985. Effect of suspension potential on the oxidation rate of copper
(Hamburg, Germany). Hydrometallurgy, vol. 5. GDMB, Clausthal-Zellerfeld, concentrate in a sulfuric acid solution. Metall. Trans. B 16, 695705.
pp. 18711897. Karavaiko, G.I., 1985. Microbiological processes for the leaching of metals from ores. In:
Fu, K.B., Lin, H., Wang, H., Wen, H.W., Wen, Z.L., 2012. Comparative study on the Torma, A.E. (Ed.), United Nations Environment Programme. USSR Commission for
passivation layers of copper sulphide minerals during bioleaching. Int. J. Miner. UNEP, Moscow, pp. 169.
Metall. Mater. 19 (10), 886892. Kariuki, S., Moore, C., McDonald, A.M., 2009. Chlorate-based oxidative hydrometallurgical
Garten, V.A., 1962. The decomposition of hydrogen peroxide by ferric ion. Aust. J. Chem. extraction of copper and zinc from copper concentrate sulde ores using mild acidic
15, 719728. conditions. Hydrometallurgy 96, 7276.
Gautier, V., Escobar, B., Vargas, T., 2008. Cooperative action of attached and planktonic Kimizuka, K., Kajiwara, S. and Tsuruga, T., 2001. Process for producing persulfate. U.S.
cells during bioleaching of chalcopyrite with Sulfolobus metallicus at 70 C. Patent 6,214,197-B1 (10 April 2001).
Hydrometallurgy 94, 121126. Kinnunen, P.H.-M., Puhakka, J.A., 2004. Chloride-promoted leaching of chalcopyrite concen-
Gericke, M., Pinches, A., 1999. Bioleaching of copper sulphide concentrate using extreme trate by biologically-produced ferric sulfate. J. Chem. Technol. Biotechnol. 79, 830834.
thermophilic bacteria. Miner. Eng. 12, 893904. Klauber, C., 2003. Fracture-induced reconstruction of a chalcopyrite (CuFeS2) surface.
Gericke, M., Pinches, A., van Rooyen, J.V., 2001. Bioleaching of a chalcopyrite concentrate Surf. Interface Anal. 35, 415428 (and Klauber, C., 2003. Erratum, op. cit. 35, 770).
using an extremely thermophilic culture. Int. J. Miner. Process. 62, 243255. Klauber, C., 2008. A critical review of the surface chemistry of acidic ferric sulphate dissolu-
Gericke, M., Neale, J.W., van Staden, P.J., 2009. A Mintek perspective of the past 25 years in tion of chalcopyrite with regards to hindered dissolution. Int. J. Miner. Process. 86, 117.
minerals bioleaching. J. South. Afr. Inst. Min. Metall. 109, 567585. Klauber, C., Parker, A., van Bronswijk, W., Watling, H.R., 2001. Sulphur speciation of
Gericke, M., Govender, Y., Pinches, A., 2010. Tank bioleaching of low-grade chalcopyrite leached chalcopyrite surfaces as determined by X-ray photoelectron spectroscopy.
concentrates using redox control. Hydrometallurgy 104, 414419. Int. J. Miner. Process. 62, 6594.
Gmez, C., Figueroa, M., Muoz, J., Blzquez, M.L., Ballester, A., 1996. Electrochemistry of Kohr, W.J., Shrader, V., Johansson, C., 2004. High temperature heap bioleaching process.
chalcopyrite. Hydrometallurgy 43, 331344. U.S. Patent 6,8028,88 B2 (12 October 2004).
Gmez, E., Ballester, A., Gonzlez, F., Blzquez, M.L., 1999. Leaching capacity of a new Kohr, W.J., Shrader, V. and Johansson, C., 2011. High temperature heap leaching process.
extremely thermophilic microorganism, Sulfolobus rivotincti. Hydrometallurgy 52, U.S. Patent 8,012,238 (6 September 2011).
349366. Kolodziej, B., Adamski, Z., 1984. A ferric chloride hydrometallurgical process for recovery
Habashi, F., 1973. The action of nitric acid on chalcopyrite. Trans. AIME 254, 224228. of silver from electronic scrap materials. Hydrometallurgy 12, 117127.
Habashi, F., 1999. Nitric acid in the hydrometallurgy of suldes. In: Mishra, B. (Ed.), Kolthoff, I.M., Meehan, E.J., Kimura, M., 1973. Hydrogen peroxide formation upon
Proceedings of the EPD Congress 1999. TMSAIME, Warrendale, PA, pp. 2532. oxidation of oxalic acid in presence and absence of oxygen and of manganese(II)
Hackl, R.P., Dreisinger, D.B., Peters, E., King, J.A., 1995. Passivation of chalcopyrite during III. Iron(III) and copper(II) as retarders. Talanta 20, 8187.
oxidative leaching in sulfate media. Hydrometallurgy 39, 2548. Konishi, Y., Asai, S., Tokushige, M., Suzuki, T., 1999. Kinetics of the bioleaching of
Halinen, A.K., Rahunen, N., Kaksonen, A.H., Puhakka, J.A., 2009. Heap bioleaching of a chalcopyrite concentrate by acidophilic thermophile Acidianus brierleyi. Biotechnol.
complex sulde ore. Part I: effect of pH on metal extraction and microbial Prog. 15, 681688.
composition in pH controlled columns. Hydrometallurgy 98, 92100. Konishi, Y., Tokushige, M., Asai, S., Suzuki, T., 2001. Copper recovery from chalcopyrite
Harvey, T.J., Bath, M., 2007. The GeoBiotics GEOCOAT technology progress and challenges. concentrate by acidophilic thermophile Acidianus brierleyi in batch and continuous-
In: Rawlings, D.E., Johnson, D.B. (Eds.), Biomining. Springer, Berlin, pp. 97112. ow stirred tank reactors. Hydrometallurgy 59, 271282.
Havlik, T., Skrobian, M., 1990. Acid leaching of chalcopyrite in the presence of ozone. Can. Lastra, M.R. and Budden, J.R., 2002. Improved high yield bioheap leaching of chalcopyrite
Metall. Q. 29, 133139. copper ores. World Patent WO 02/070758 A1 (12 September 2002).
Havlik, T., Dvorscikova, J., Ivanova, Z., Kammel, R., 1999. Sulphuric acid chalcopyrite Li, A., Huang, S., 2011. Comparison of the electrochemical mechanism of chalcopyrite
leaching using ozone as oxidant. Mtallurgie (Berl.) 53, 5760. dissolution in the absence or presence of Sulfolobus metallicus at 70 C. Miner. Eng.
Herreros, O., Quiroz, R., Longueira, H., Fuentes, G., Vials, J., 2006. Leaching of djurleite in 24, 15201522.
Cu2+/Cl media. Hydrometallurgy 82, 3239. Li, D., Li, D., 2010. Citric catalysis on the microbiological leaching of chalcopyrite ne
Hiroyoshi, N., Nakamura, T., Tsunekawa, M., Hirajima, T., Ito, M., 1995. Enhancement in grained tailings in shake ask. Res. J. Biotechnol. 5, 6467.
bacterial leaching of chalcopyrite by polyoxyethylene sorbitan monolaurate addition. Li, J., Kawashima, N., Kaplun, K., Absolon, V.J., Gerson, A.J., 2010. Chalcopyrite leaching: the
J. Min. Mater. Process. (Jpn) 111, 943948. rate controlling factors. Geochim. Cosmochim. Acta 74, 28812893.
Hiroyoshi, N., Hirota, M., Hirajima, T., Tsunekawa, M., 1997. A case of ferrous sulfate Li, Y., Kawashima, N., Li, J., Chandra, A.P., Gerson, A.P., 2013. A review of the structure and
addition enhancing chalcopyrite leaching. Hydrometallurgy 47, 3745. fundamental mechanisms and kinetics of the leaching of chalcopyrite. Adv. Colloid
Hiroyoshi, N., Miki, H., Hirajima, T., Tsunekawa, M., 2000. A model for ferrous-promoted Interf. Sci. 197198, 132.
chalcopyrite leaching. Hydrometallurgy 57, 3138. Liang, C.L., Xia, J.L., Nie, Z.Y., Yang, Y., Ma, C.Y., 2012. Effect of sodium chloride on sulfur
Hiroyoshi, N., Arai, M., Miki, H., Tsunekawa, M., Hirajima, T., 2002. A new reaction model speciation of chalcopyrite bioleached by the extreme thermophile Acidianus
for the catalytic effect of silver ions on chalcopyrite leaching in sulfuric acid solutions. manzaensis. Bioresour. Technol. 110, 462467.
Hydrometallurgy 63, 257267. Lin, H.K., Luong, H.V., 2004. Column leaching for simulating heap and in-situ soil
Hiroyoshi, N., Kuroiwa, S., Miki, H., Tsunekawa, M., Hirajima, T., 2007. Effects of coexisting remediation with metallic Fenton reaction. J. Miner. Mater. Charact. Eng. 3, 3339.
metal ions on the redox potential dependence of chalcopyrite leaching in sulfuric acid Linge, H.G., 1976. A study of chalcopyrite dissolution in acidic ferric nitrate by
solutions. Hydrometallurgy 87, 110. potentiometric titration. Hydrometallurgy 2, 5164.
Hiroyoshi, N., Kitagawa, H., Tsunekawa, M., 2008. Effect of solution composition on the Lu, Z.Y., Jeffrey, M.I., Lawson, F., 2000. The effect of chloride ions on the dissolution of
optimum redox potential for chalcopyrite leaching in sulfuric acid solutions. chalcopyrite in acidic solutions. Hydrometallurgy 56, 189202.
Hydrometallurgy 91, 144149. Lunt, D., Poulter, S., Canterford, J., 1997. Hydrometallurgical treatment of copper
Hourn, M., Halbe, D., 1999. The NENATECH Process: Results on Frieda River copper gold concentrates: what are the real process options? ALTA 1997 Copper Hydro-
concentrates. Randol Copper Hydromet Roundtable. Randol International, Golden metallurgy Forum (Brisbane). ALTA Metallurgical Services, Melbourne (24 pp.).
Colorado, pp. 97102. Mackie, D., Trask, F., 2009. Continuous vat leaching rst copper pilot trials. ALTA
Hourn, M.M., Turner, D.W., Holzberger, I.R., 1999. Atmospheric mineral leaching process. Copper-X. ALTA Metallurgical Services, Melbourne.
U.S. Patent 5,993,635 (30 November 1999). Mahajan, V., Misra, M., Zhong, K., Fuerstenau, M.C., 2007. Enhanced leaching of copper
Hu, Y., Qiu, G., Wang, J., Wang, D., 2002. The effect of silver-bearing catalysts on from chalcopyrite in hydrogen peroxideglycol system. Miner. Eng. 20, 670674.
bioleaching of chalcopyrite. Hydrometallurgy 64, 8188. Mallick, S.A., Dasgupta, A.K., 1997. Bioleaching of copper from chalcopyrite ores coated
Hunter, C.J., 2002a. Bioheap leaching of a primary nickelcopper sulphide ore. ALTA with polyaniline lm. Curr. Sci. 72, 201205.
Nickel/Cobalt 2002 (Perth). ALTA Metallurgical Services, Melbourne (11 pp.). Mazuelos, A., Iglesias, N., Carranza, F., 1999. Inhibition of bioleaching processes by
Hunter, C.J., 2002b. Method for the bacterially assisted heap leaching of chalcopyrite. organics from solvent extraction. Process Biochem. 35, 425431.
World Patent 02/070757 A1 (12 September 2002). McDonald, R.G., Muir, D.M., 2007a. Pressure oxidation leaching of chalcopyrite. Part 1.
Hunter, C.J., 2006. Method for the bacterially assisted heap leaching of chalcopyrite. U.S. Comparison of high and low temperature reaction kinetics and products.
Patent 7,022,504 (4 April 2006). Hydrometallurgy 86, 191205.
Hunter, C.J., Williams, T.L., 2006. Heap leach. U.S. Patent 2006/0248983 A1 (9 November McDonald, R.G., Muir, D.M., 2007b. Pressure oxidation leaching of chalcopyrite. Part 2.
2006). Comparison of medium temperature kinetics and products and effect of chloride
Index Mundi, 2013. Copper: World Production by Country. Source: www.indexmundi. ion. Hydrometallurgy 86, 206220.
com/en/commodities/minerals/copper/copper_t20.html. Mier, J.L., Gomez, C., Ballester, A., Blazquez, M.L., Gonzalez, F., 1994. The effects of silver
International Copper Study Group, 2012. World Copper Factbook 2012. 54 (Source: and bismuth on bioleaching of copper sulphide concentrates with thermophilic
www.icsg.org/index.php/component/jdownloads/nish/170/1188). microorganisms. Hydrometallurgy '94. IMM, Chapman and Hall, London, pp. 369383.
180 H.R. Watling / Hydrometallurgy 140 (2013) 163180

Miller, J.D., Portillo, H.Q., 1979. Silver catalysis in ferric sulfate leaching of chalcopyrite. In: Ruan, R., Zhou, E., Liu, X., Wu, B., Zhou, G., Wen, J., 2010. Comparison on the leaching
Laskowski, J. (Ed.), Proceedings of the XIII Mineral Processing Congress. Elsevier, kinetics of chalcopyrite and pyrite with or without bacteria. Rare Metals 29, 552556.
Amsterdam, pp. 851901. Ruiz, M.C., Montes, K.S., Padilla, R., 2011. Chalcopyrite leaching in sulfatechloride media
Mudd, G.M., 2010. The environmental sustainability of mining in Australia: key mega- at ambient pressure. Hydrometallurgy 109, 3742.
trends and looming constraints. Resour. Policy 35, 98115. Salo-Zieman, V.L.A., Kinnunen, P.H.M., Puhakka, J.A., 2006. Bioleaching of acid-consuming
Munoz, P.B., Miller, J.D., Wadsworth, M.E., 1979. Reaction mechanism for the acid ferric low-grade nickel ore with elemental sulfur addition and subsequent acid generation.
sulfate leaching of chalcopyrite. Metall. Trans. B 10, 149158. J. Chem. Technol. Biotechnol. 81, 3440.
Muoz, J.A., Gmez, C., Ballester, A., Blazquez, M.L., Gonzalez, F., 1998. Electrochemical Sandstrm, ., Petersson, S., 1997. Bioleaching of a complex sulphide ore with moderate
behaviour of chalcopyrite in the presence of silver and Sulfolobus bacteria. J. Appl. thermophilic and extreme thermophilic microorganisms. Hydrometallurgy 46, 181190.
Electrochem. 28, 4956. Sandstrm, ., Shchukarev, A., Paul, J., 2005. XPS characterisation of chalcopyrite
Muoz, J.A., Dreisinger, D.B., Cooper, W.C., Young, S.K., 2007a. Silver catalyzed bioleaching chemically and bio-leached at high and low redox potential. Miner. Eng. 18, 505515.
of low-grade copper ores. Part I: shake ask tests. Hydrometallurgy 88, 318. Sawyer, H.D., Shaw, R.W., 1983. Hydrometallurgical recovery of copper. Australian Patent
Muoz, J.A., Dreisinger, D.B., Cooper, W.C., Young, S.K., 2007b. Silver catalysed bioleaching Au-B-17181/83.
of low-grade copper ores. Part III. Column reactors. Hydrometallurgy 88, 3551. Schlesinger, M.E., King, M.J., Sole, K.C., Davenport, W.G.I., 2011. Hydrometallurgical copper
Muoz-Ribadeneira, F.J., Gomberg, H.J., 1970. Possibilities of recovering copper from extraction: Introduction and leaching, In: Schlesinger, M.E., King, M.J., Sole, K.C.,
chalcopyrite by underground leaching in situ. Trans. Am. Nucl. Soc. 13, 636637. Davenport, W.G. (Eds.), Extractive Metallurgy of Copper, Fifth ed. Elsevier, Oxford,
Muoz-Ribadeneira, F.J., Gomberg, H.J., 1971. Leaching of chalcopyrite (CuFeS2) with pp. 281322.
sodium chloride sulfuric acid solutions. Nucl. Technol. 11 (3), 367371. Schlitt, W.J., 2006. Kennecott's million-ton test heap the active leach program. Miner.
Murr, L.E., Hiskey, G.B., 1981. Kinetic effect of particle size and crystal dislocation density Metall. Process. 23 (1), 116.
on the dichromate leaching of chalcopyrite. Metall. Trans. B 12, 255267. Schlitt, J., Johnston, A., 2010. The Marcobre vat leach system: a new look at an old process.
Naderi, H., Abdollahy, M., Mostou, N., Koleini, M.J., Shojaosadati, S.A., Mana, Z., 2011. Proceedings of Copper 2010. Hydrometallurgy, vol. 5. GDMB, Clausthal-Zellerfeld,
Kinetics of chemical leaching of chalcopyrite from low grade copper ore: behaviour pp. 20392057.
of different size fractions. Int. J. Miner. Metall. Mater. 18, 638645. Schnell, H., 1997. Bioleaching of copper. In: Rawlings, D.E. (Ed.), Biomining: Theory,
Nazari, G., Dixon, D.G., Dreisinger, D.B., 2012. The mechanism of chalcopyrite leaching in Microbes and Industrial Processes. Springer Verlag, Berlin, pp. 2143.
the presence of silver-enhanced pyrite in the Galvanox process. Hydrometallurgy Shantz, R., Morris, T.M., 1974. Dichromate process demonstrated for leaching of copper
113114, 122130. sulphide concentrates. Eng. Min. J. 175, 7172.
Nicol, M.J., Lzaro, I., 2003. The role of non-oxidative processes in the leaching of Sokic, M.D., Markovic, B., Zivkovic, D., 2009. Kinetics of chalcopyrite leaching by sodium
chalcopyrite. In: Riveros, P.A., Dixon, D., Dreisinger, D.B., Menacho, J. (Eds.), Copper nitrate in sulphuric acid. Hydrometallurgy 95, 273279.
2003Cobre 2003, vol. VI. Canadian Institute of Mining, Metallurgy and Petroleum, Sokic, M.D., Matkovic, V.L., Markovic, B.R., Strbac, N.D., Zivkovic, D.T., 2010. Passivation of
Montreal, pp. 367381 (Book 1). chalcopyrite during the leaching with sulphuric acid solution in presence of sodium
O'Brien, R.T., MacDonald, C.A., Meadows, N.E., 1999. Chloridesulphate leaching from nitrate. Hem. Ind. 64, 343350.
chalcopyrite ores from Sabah. ALTA Copper Sulphides Symposium and Copper Solis-Marcal, O.J., Lapidus, G.T., 2013. Improvement of chalcopyrite dissolution in acid
Hydrometallurgy Forum (Gold Coast, QLD). ALTA Metallurgical Services, Melbourne. media using polar organic solvents. Hydrometallurgy 131132, 120126.
Okibe, N. And, Johnson, D.B., 2002. Toxicity of otation reagents to moderately Stott, M.B., Sutton, D.C., Watling, H.R., Franzmann, P.D., 2001. The effect of solution
thermophilic bioleaching microorganisms. Biotechnol. Lett. 24, 20112016. chemistry on jarosite deposition during the leaching of chalcopyrite by the
Olubambi, P.A., Potgieter, J.H., 2009. Investigations on the mechanisms of sulfuric acid thermophilic archaeon Sulfolobus metallicus. Process Metall. 11A, 207215.
leaching of chalcopyrite in the presence of hydrogen peroxide. Miner. Process. Extr. Subramanian, K.N., Ferrajuolo, R., 1976. Oxygen pressure leaching of FeNiCu sulde
Metall. Rev. 30, 327345. concentrates at 110 C effect of low chloride addition. Hydrometallurgy 2, 117126.
Olubambi, P.A., Borode, J.O., Ndlovu, S., 2006. Sulphuric acid leaching of zinc and copper Third, K.A., Cord-Ruwisch, R., Watling, H.R., 2002. Control of redox potential by oxygen
from Nigerian complex sulphide ore in the presence of hydrogen peroxide. J. S. Afr. limitation improves bacterial leaching of chalcopyrite. Biotechnol. Bioeng. 78, 433441.
Inst. Min. Metall. 106, 765770. Torma, A.E., Itzkovitch, I.J., 1976. Inuence of organic solvents on chalcopyrite oxidation
Palencia, I., Romero, R., Carranza, E., 1998. Silver-catalyzed IBES process: application to a ability of Thiobacillus ferrooxidans. Appl. Environ. Microbiol. 32, 102107.
Spanish copperzinc sulphide concentrate. Part 2. Biooxidation of the ferrous iron Uhrie, J.L., Bowman, G., Caro Matus, C.A., Mayta, P., Hoenecke, S., Chavez, M., 2012.
and catalyst recovery. Hydrometallurgy 48, 101112. Methods and systems for leaching a metal-bearing ore for the recovery of a metal
Palmer, B.R., Nebo, C.O., Rau, M.F., Fuerstenau, M.C., 1981. Rate phenomena involved in value. U.S. Patent 8,118,907 (21 February, 2012).
the dissolution of chalcopyrite in chloride-bearing lixiviants. Metall. Trans. B 12, van der Merwe, C., Pinches, A., Myburgh, P.J., 1998. A process for the leaching of
595601. chalcopyrite. World Patent WO9839491 (11 September 1998); also published as
Parker, A.J., Paul, R.L., Power, G.P., 1981. Electrochemistry of the oxidative leaching of Australian Patent 2224697 and subsequently as Pinches et al., 2001b.
copper from chalcopyrite. J. Electroanal. Chem. Interfacial Electrochem. 118, 305316. van Staden, P.J., 1998. The Mintek/Bactech copper bioleach process. ALTA 1998 Copper
Parker, A., Klauber, C., Kougianos, A., Watling, H.R., van Bronswijk, W., 2003. An X-ray Sulphides Symposium (Brisbane). ALTA Metallurgical Services, Melbourne.
photoelectron spectroscopy study of the mechanism of oxidative dissolution of van Staden, P.J., Shaidaee, B., Yazdani, M., 2005. A collaborative plan towards heap
chalcopyrite. Hydrometallurgy 71, 265276. bioleaching of low grade chalcopyritic ore from a new Iranian mine. In: Harrison,
Peacey, J., Guo, X., Robles, E., 2004. Copper hydrometallurgy current status, preliminary S.T.L., Rawlings, D.E., Petersen, J. (Eds.), Proceedings of the 16th International
economics, future direction and positioning versus smelting. Trans. Nonferrous Met. Biohydrometallurgy Symposium (Cape Town), 16th International Biohydrometallurgy
Soc. China 14, 560568. Symposium, Cape Town, pp. 115123.
Philippe, R., Dixon, R., Dal Pozzo, S., 2010. Sal or Desal. Seawater supply options for the mining Viramontes-Gamboa, G., Rivera-Vasquez, B.F., Dixon, D.G., 2007. The activepassive
industry. Paper presented at water in mining. II International Congress on Water behavior of chalcopyrite comparative study between electrochemical and leaching
Management in the Mining Industry (Santiago, Chile). Gecamin, Santiago, pp. 1320. responses. J. Electrochem. Soc. 154, C299C311.
Pinches, A., Myburgh, P.J., van der Merwe, C., 2001b. Process for the rapid leaching of Viramontes-Gamboa, G., Pea-Gomar, M.M., Dixon, D.G., 2010. Electrochemical hysteresis
chalcopyrite in the absence of catalysts. U.S. Patent 6,277,341 (21 August 2001). and bistability in chalcopyrite passivation. Hydrometallurgy 105, 140147.
Pinches, A., Gericke, M., and van Rooyen, J.V., 2001a. A method of operating a bioleach Vraar, R.Z., Parezanovi, I.S., Cerovi, K.P., 2000. Leaching of copper(I) sulde in chloride
process with control of redox potential. World Patent 01/31072 A1 (3 May 2001). solution. Hydrometallurgy 58, 261267.
Pradhan, N., Nathsarma, K.C., Srinivasa Rao, K., Sukla, L.B., Mishra, B.K., 2008. Heap Vraar, R.Z., Vuckovi, N., Kamerovi, Z., 2003. Leaching of copper(I) sulphide by sulphuric
bioleaching of chalcopyrite: a review. Miner. Eng. 21, 355365. acid solution with addition of sodium nitrate. Hydrometallurgy 70, 143151.
Prater, J.D., Queneau, P.B., Hudson, T.J., 1973. Nitric acid route to processing copper Wang, S., 2005. Copper leaching from chalcopyrite concentrates. J. Met. 57 (7), 4851.
concentrate. Trans. AIME 254, 117122. Warren, G.W., Wadsworth, M.E., El-Raghy, S.M., 1982. Passive and transpassive anodic
Price, D.W., Warren, G.W., 1986. The inuence of silver ion on the electrochemical behaviour of chalcopyrite in acid solutions. Metall. Trans. B 13, 571579.
response of chalcopyrite and other mineral sulde electrodes in sulfuric acid. Warren, G.W., Drouven, B., Price, D.W., 1984. Relationships between the Pourbaix
Hydrometallurgy 15, 303324. diagram for AgSH2O and electrochemical oxidation and reduction of Ag2S. Metall.
Queneau, P.B. and Prater, J.D., 1974. Nitric acid process for recovering metal values from Trans. B 15, 235242.
sulde ore materials containing iron sulde. U.S. Patent 3,793,429 (19 February 1974). Watling, H.R., 2006. The bioleaching of sulphide minerals with emphasis on copper
Radimer, K.J. and McCarthy, M.J., 1979. Electrolytic production of sodium persulfate. U.S. sulphides a review. Hydrometallurgy 84, 81108.
Patent 4,144,144 (13 March 1979). Watling, H.R., Elliot, A.D., Maley, M., van Bronswijk, W., Hunter, C., 2009. Leaching of a
Rauld Faine, J., Aroca Alfaro, F., Montealegre Jullian, R., Backit Gutierrez, A., 2005a. low-grade, coppernickel sulde ore. 1. Key parameters impacting on Cu recovery
Non-biochemical method to heap leach copper concentrates. U.S. Patent 6,926,753 during column bioleaching. Hydrometallurgy 97, 204212.
(9 August 2005). Westby, G.C., 1918. Nitric acid and copper ores. Metall. Chem. Eng. 18, 290296.
Rauld Faine, J., Aroca Alfaro, F., Montealegre Jullian, R., Backit Gutierrez, A., 2005b. Xia, L., Tang, L., Xia, J., Yin, C., Cai, L., Zhao, X., Nie, Z., Liu, J., Qiu, G., 2012. Relationships
Procedure to leach copper concentrates, under pressure and at ambient temperature, among bioleaching performance, additional sulfur, microbial population dynamics
by forming a reactive gel in a sulfatechloride medium. US Patent 2005/0169823 A1 and its energy metabolism in bioleaching of chalcopyrite. Trans. Nonferrous Met.
(process also described in European Patent EP 1559799). Soc. China 22, 192198.
Rodrguez, Y., Ballester, A., Blzquez, M.L., Gonzlez, F., Muoz, J.A., 2003. New in- Xian, Y.J., Wen, S.M., Deng, J.S., Liu, J., Nie, Q., 2012. Leaching chalcopyrite with sodium
formation on the chalcopyrite bioleaching mechanism at low and high temperature. chlorate in hydrochloric acid solution. Can. Metall. Q. 52, 133140.
Hydrometallurgy 71, 4756. Young, T.L., Greene, M.G., Rice, D.R., Karlage, K.L., Premeau, S.P., Cassells, J.M., 2003.
Romero, R., Mazuelos, A., Palencia, I., Carranza, F., 2003. Copper recovery from Method for the microbiological production of sulfuric acid. U.S. Patent 6,610,268
chalcopyrite concentrates by the BRISA process. Hydrometallurgy 70, 205215. (26 August 2003).

S-ar putea să vă placă și