Sie sind auf Seite 1von 23

Green

View Article Online


View Journal

Chemistry
Cutting-edge research for a greener sustainable future

Accepted Manuscript

This article can be cited before page numbers have been issued, to do this please use: S. Anggara, F.
Bevan, R. C. Harris, J. Hartley, G. Frisch, G. Jenkin and A. P. Abbott, Green Chem., 2019, DOI:
10.1039/C9GC03213D.
Volume 20
This is an Accepted Manuscript, which has been through the
Green
Number 9
7 May 2018
Pages 1919-2160
Royal Society of Chemistry peer review process and has been
accepted for publication.
Chemistry
Cutting-edge research for a greener sustainable future Accepted Manuscripts are published online shortly after acceptance,
rsc.li/greenchem

before technical editing, formatting and proof reading. Using this free
service, authors can make their results available to the community, in
citable form, before we publish the edited article. We will replace this
Accepted Manuscript with the edited and formatted Advance Article as
soon as it is available.

You can find more information about Accepted Manuscripts in the


Information for Authors.

Please note that technical editing may introduce minor changes to the
text and/or graphics, which may alter content. The journal’s standard
ISSN 1463-9262 Terms & Conditions and the Ethical guidelines still apply. In no event
PAPER
Paul T. Anastas et al.
The Green ChemisTREE: 20 years after taking root with the 12
shall the Royal Society of Chemistry be held responsible for any errors
principles

or omissions in this Accepted Manuscript or any consequences arising


from the use of any information it contains.

rsc.li/greenchem
Page 1 of 22 Green Chemistry

Direct Extraction of Copper from Copper Sulfide Minerals using


DOI:Deep
View Article Online
10.1039/C9GC03213D

Eutectic Solvents
Syahrie Anggara,a Francesca Bevan,a Robert C. Harris,a Jennifer M. Hartley,a,b Gero Frisch,b
Published on 08 November 2019. Downloaded by Centro de Desenvolvimento da Tecnologia Nuclear on 11/11/2019 8:14:34 PM.

Gawen R. T. Jenkin,c Andrew P. Abbotta*

a Department of Chemistry, University of Leicester, Leicester, LE1 7RH, UK

Green Chemistry Accepted Manuscript


b Institut für Anorganische Chemie, TU Bergakademie Freiberg, 09599 Freiberg, Germany
c School of Geography, Geology and the Environment, University of Leicester, Leicester,
LE1 7RH, UK

E-mail: apa1@le.ac.uk;

Abstract
Copper is predominantly recovered from sulfide ores by pyrometallurgy – an energy intensive
process requiring capture and treatment of released SO2. Whilst some copper ores are amenable
to hydrometallurgy, chalcopyrite, the main copper mineral, is challenging to process in aqueous
solutions due to surface passivation. The chalcopyrite surface may be less prone to passivation
in non-aqueous solvents such as Ionic Liquids. Here we provide the first demonstration that
electrochemistry in Deep Eutectic Solvents can solubilise and directly recover high purity
copper from solution from three copper sulfide minerals: covellite (CuS), chalcocite (Cu2S)
and chalcopyrite (CuFeS2). Cyclic voltammetry supported by EXAFS identifies the metal
speciation in solution. In a choline chloride-ethylene glycol DES the main copper species
present after dissolution of chalcocite and covellite was [CuCl4]2−. In the solution formed from
chalcopyrite, a mix of CuII and CuI species were formed instead. In a choline chloride-urea
DES, copper had a mixed chloride/O- or N-donor coordination, potentially altering
electrochemical behaviour. Sulfide in the mineral is oxidised to sulfate without the generation
of SO2 or H2S. The best selective recovery of copper (99 at%) from chalcopyrite was obtained
with a mixed DES of 20 wt% choline chloride-oxalic acid and 80 wt% choline chloride-
ethylene glycol. This demonstrates how design of the Deep Eutectic Solvents can enable
increased selectivity of copper over iron in the electrowinning stage by changing solute
speciation and redox properties.
Key words
Chalcopyrite, paint casting, copper, sulfide, deep eutectic solvent, electrowinning, EXAFS

1
Green Chemistry Page 2 of 22

View Article Online


DOI: 10.1039/C9GC03213D

Introduction
Copper ores are most commonly processed through pyrometallurgical methods, and currently
around 84% of copper (nearly 17 million tonnes per annum) is produced this way, dominantly
Published on 08 November 2019. Downloaded by Centro de Desenvolvimento da Tecnologia Nuclear on 11/11/2019 8:14:34 PM.

from chalcopyrite ores that are not amenable to hydrometallurgy.1-3 Ores with significant
impurities, such as >1% arsenic, present environmental and technical challenges for smelters
and alternative process methods are desirable, i.e. hydrometallurgical methods which use

Green Chemistry Accepted Manuscript


aqueous solutions to solubilise metal species,4, 5 after which the metals are subsequently
recovered by methods such as precipitation, electrowinning, cementation, or ion exchange.2
However, these traditional methods for the processing of metals often have high energy usages
and/or utilise corrosive solutions, which can contain high concentrations of acid or base, and
can involve the production of acidic gases such as H2S and SO2, especially where sulfide
minerals are the main feedstock.6

The secondary-enriched oxide and sulfide zones of copper deposits that are formed near-
surface in the weathering zone are characterised by copper oxides, carbonates, and sulfide
minerals such as chalcocite (Cu2S) and covellite (CuS) that are easily and economically
processed by heap leaching with sulfuric acid. However, as these easily exploitable near-
surface deposits become exhausted, the deeper lower grade primary ores dominated by
chalcopyrite (CuFeS2) are increasingly being exploited. Chalcopyrite, which accounts for
approximately 70% of world copper reserves,7 is extremely difficult to leach under
hydrometallurgical conditions. It has been suggested that slow dissolution rates are due to
passivation of the mineral surface via growth of an impermeable sulfur or polysulfide layer.8, 9
Aqueous leachates for chalcopyrite can require high temperatures,10 additives such as
ferrous/ferric species11, 12 and peroxides,13 although more environmentally friendly methods
such as microbial leaching have also been investigated.1, 14

In an alternative drive to improve the environmental hazards of classical acidic and caustic
hydrometallurgical processing, ionometallurgy has been proposed as an alternative approach.
Ionic liquids (ILs) enable high solubility of metals salts and wide electrochemical windows,
which are useful properties for mineral processing.15 Interestingly, it has been observed that
the presence of chloride ions in aqueous solutions can increase the rate of chalcopyrite leaching
due to the formation of chlorocuprate(I) ions,16 and a key property of ILs is their high anionic
component content. As the anionic component of the IL generally controls speciation of metals,
2
Page 3 of 22 Green Chemistry

the need for highly coordinating additives is reduced, potentially allowing novelDOI:
dissolution
View Article Online
10.1039/C9GC03213D

routes. Investigations into the leaching of chalcopyrite in imidazolium ionic liquids indicate a
highly complex dissolution mechanism, with sulfide and polysulfide species forming in many
liquids.17 The presence of FeIII species in IL media has an analogous effect to aqueous media,
Published on 08 November 2019. Downloaded by Centro de Desenvolvimento da Tecnologia Nuclear on 11/11/2019 8:14:34 PM.

with the initial rate of chalcopyrite leaching being improved compared to a FeIII-free system.18
[C4mim][HSO4] has been suggested as an alternative to H2SO4 as a leachate for chalcopyrite,
due to its ability to catalyse the transfer of oxygen and accelerate the oxidation of

Green Chemistry Accepted Manuscript


chalcopyrite,19, 20 with the potential for solvent recyclability.21

It has been reported by Carlesi et al.22 that acidic ILs containing the [HSO4]‒ anion can promote
chalcopyrite dissolution in combination with 1 M H2SO4. It was observed that [C6mim][HSO4]
and [C4mim][HSO4] both exhibit some capacity to dissolve chalcopyrite at higher
temperatures, however there were no clear advantages over simple leaching with H2SO4. It was
observed, however, that when these ILs were used in combination with 1 M H2SO4 in equal
volumetric amounts, copper recovery rates were greatly improved (~70%) in comparison to the
aqueous leaching (~24%) when the temperatures exceeded 60 °C. It was proposed that the main
role of the IL is to enable continuous acid leaching by promoting the decomposition of potential
sulfidic intermediates that can passivate the surface of the chalcopyrite. Carlesi also suggests
that the hydrophobicity of the IL is important in removing oxidation products that are insoluble
in aqueous media from the surface of the chalcopyrite. This has been supported by Albrecht et
al.23 who tested a range of HSO4‒ ILs and found that the hydrophobicity of the IL plays a key
role in its effectiveness as a potential lixiviant for chalcopyrite. It was observed in this study
that ILs [C1mim][HSO4] and [NH4][HSO4] were the best performing and demonstrated that the
higher the polarity the cation is, the poorer the wetting of the surface of the chalcopyrite. This
in turn resulted in higher Cu extraction rates than the less polar ILs investigated. Hu et al.24
demonstrated that ILs combined with hydrogen peroxide (H2O2) as an oxidant can yield high
yields (>98%) of Cu from chalcopyrite in 2 h at 45°C. The use of ILs in commercial
applications is however severely limited by their relatively high cost and so full scale
applications using this technology become very difficult to justify.

Chloride leaching of chalcopyrite has been shown to be a potential alternative to the majority
of sulfate-based lixiviants, which often exhibit low kinetic rates of dissolution of
chalcopyrite.25, 26 Early work had demonstrated that chloride-based lixiviants such as aqueous
CuCl2 and FeCl3 could recover >95% of Cu from chalcopyrite. A mixed chloride-sulfate bath
3
Green Chemistry Page 4 of 22

was demonstrated in work carried out by Jeffrey et al.,27 who showed that adding DOI:
0.510.1039/C9GC03213D
M NaCl
View Article Online

to a 1 M H2SO4 solution significantly reduced the rate of passivation of chalcopyrite during


electrochemical oxidation. Fuentes-Aceituno et al.28 demonstrated that electro-assisted
dissolution of chalcopyrite in a 2 M HCl solution exhibits a greater efficiency of chalcopyrite
Published on 08 November 2019. Downloaded by Centro de Desenvolvimento da Tecnologia Nuclear on 11/11/2019 8:14:34 PM.

dissolution than in analogous H2SO4 solutions. In both cases this improved dissolution rate of
chalcopyrite in the chloride-containing systems is due to the destabilising effect chloride has
on the sulfide-based passivation layer that forms during the dissolution of chalcopyrite.

Green Chemistry Accepted Manuscript


Dresinger et al.29 devised a number of process flowsheets using solutions of either CuCl2·2H2O
or FeCl3 that resulted in Fe-free Cu solutions, which could then be transferred into a
conventional sulfate-based electrowinning circuit. In both processes >95% of the Cu in
chalcopyrite could be extracted, with dissolved Fe precipitated out as hematite (Fe2O3) or
goethite (FeOOH).

Deep Eutectic Solvents (DESs) are primarily chloride based media that offer similar properties
to ILs, however they are cheaper and easier to produce on a large scale and can be designed to
have low toxicity.30 DESs have been demonstrated to be effective in the dissolution and
extraction of other metals, for example in the oxidation of gold31, 32 and several sulfides,33, 34
rare earth carbonates,35 or in the processing of electric arc furnace36 and flue dusts.37 The
current study investigates the electrochemical behaviour of three common copper sulfide
minerals in DES media. Chalcopyrite is composed of monovalent copper with trivalent iron in
a monosulfide: CuIFeIIIS2.38 Covellite and chalcocite are both also thought to contain
monovalent copper, but chalcocite is a simple monosulfide, whereas covellite is a mixed
monosulfide and disulphide mineral.39, 40 Their electrochemistry is explored using a paint
casting method, and the speciation of the dissolved metal ions is characterised via spectroscopic
methods. A novel bulk digestion and recovery process for high purity copper from natural
chalcopyrite is demonstrated, whilst avoiding hazardous by-products such as H2S or SO2.
Through the use of hybrid DESs, we highlight the potential of ionometallurgy for the
processing of copper ores by judicious tuning of the DES composition to obtain a high purity
of electrowon copper.

Experimental
All DESs were made using the methods described in previous publications.41, 42 Mixtures were
made with 1 mol. eq. of choline chloride (ChCl) and 2 mol eq. of ethylene glycol (EG), urea
(U), or 1 mol equivalent of oxalic acid dihydrate (Ox), hereafter referred to as 1ChCl:2EG,
4
Page 5 of 22 Green Chemistry

1ChCl:2U, and 1ChCl:Ox, respectively. All chemicals were obtained from Aldrich, with
View Article Online
DOI: 10.1039/C9GC03213D

purities of > 99%.

Natural covellite (sample 20582 from Butte Montana) and chalcocite (non-curated sample)
Published on 08 November 2019. Downloaded by Centro de Desenvolvimento da Tecnologia Nuclear on 11/11/2019 8:14:34 PM.

were obtained from the School of Geography, Geology and the Environment at the University
of Leicester, whilst natural chalcopyrite was sourced from Morocco. Powder X-ray diffraction
(XRD) was carried out with a Phillips model PW 1730 X-ray generator on both the synthetic

Green Chemistry Accepted Manuscript


and natural samples in order to confirm that their crystal structures are the same. Both natural
and synthetic CuS had similar XRD patterns consistent with the ICSD database entry for
covellite, confirming that the synthetic CuS has a covellite structure. The XRD pattern for
synthetic and natural chalcocite showed diffraction peaks consistent with chalcocite in the
ICSD database, but the natural sample showed some additional peaks, suggesting that there
may be slight impurities. Zeiss EDX analysis was carried out on the natural chalcopyrite to
determine mineral and impurity content of the ground and sieved material. This material had
an average composition of 80.5% chalcopyrite, with 10.7% quartz and 3.9% pyrite as the major
impurities. In decreasing order, the remaining 4.9% of the material consisted of galena,
magnetite, barite, covellite, sphalerite, clinopyroxene and chalcocite. The average elemental
composition of the chalcopyrite was close to stoichiometric, and had minimal inclusions of
other minerals.

The natural mineral samples were ground with a fly press and TEMA mill, and homogenised
via riffling. The material was then passed through a 300 µm sieve, and the fraction that did not
pass through a further 100 µm sieve was used in the paint casting experiments.

Electrochemical measurements
Cyclic voltammetry (CV) experiments were carried out using the paint casting method,
previously described by Abbott et al.,33 unless otherwise stated. The pastes which were painted
onto the electrodes were formed by finely grinding ca. 2 mg of copper mineral powder with
100 mg of DES, before being spread evenly on the surface of a Pt flag working electrode and
immersing in the electrolyte (same DES as used to make the paste).

An Autolab Type III potentiostat with GPES software was used for electrochemical
measurements. All CVs were first swept to positive potentials, starting from an initial potential
where no redox events were taking place. A three electrode system (2x Pt flag and a 100 mM
5
Green Chemistry Page 6 of 22

Ag/AgCl reference) was used for CV experiments. The electrode potentials discussed here
View Article Online
DOI: 10.1039/C9GC03213D

were estimated from averages of peak onset values. All CVs were carried out at 5 mV s-1.

For bulk dissolution and electrolysis experiments, a 2 electrode system was used. For
Published on 08 November 2019. Downloaded by Centro de Desenvolvimento da Tecnologia Nuclear on 11/11/2019 8:14:34 PM.

dissolution experiments, iridium oxide-coated titanium mesh was used as both an anode and
cathode, and the mineral paste was ‘painted’ onto the surface of the mesh before being
electrolysed for 24 h at a constant 1.5 V. In the case of the bulk electrolysis, a Ni plate was

Green Chemistry Accepted Manuscript


used as the cathode and an iridium oxide-coated titanium mesh anode was submerged into a
porous tube containing a paste of ca. 50 g of the mineral. A schematic of the bulk electrolysis
cell is depicted in Figure 1. Bulk electrolysis was carried out at a constant current density of
0.2 mA cm-2 and for each sample, the Ni cathode was replaced every 24 h. The resulting plated
deposits were analysed with a Hitachi S-3600N Environmental Scanning Electron Microscope
(SEM) with Oxford INCA 350 energy-dispersive X-ray spectroscopy (EDX) software in order
to determine their morphology and composition. The precipitates remaining at the bottom of
the cell after electrodissolution were collected, dried and weighed and their composition was
analysed with EDX.

Spectroscopic measurements
All UV-Vis spectra were recorded using an UV5Bio (Mettler Toledo) UV-Vis spectrometer
between 190 and 1000 nm. The Fe- and Cu-speciation of anodically dissolved copper minerals
were determined by extended X-ray absorption fine structure (EXAFS), at BM26A of the
ESRF synchrotron.43 Measurements were carried out at the Fe and Cu K-edges, nominally 7112
and 8979 eV, respectively. Transmission data were measured using ionisation chambers, whilst
fluorescence data were measured with a nine-element Ge solid-state detector operating on
seven elements in total fluorescence yield mode. The monochromator was a Si(111) double
crystal. Copper and iron foils were used as a reference sample for amplitude calibration. Liquid
samples were kept in a flat Perspex sample holder, with a sample chamber of 15 x 8 mm, at 1.5
mm thick, with 40 µm Kapton foil windows. The samples were aligned at approximately 55°
with respect to the X-ray beam, with the fluorescence detector placed perpendicular to the X-
ray beam in order to minimise any elastic scattering signals. Two to four spectra were recorded
for every sample, then averaged, calibrated and background subtracted with the program
Athena.44 The EXAFS spectra were fitted using EXCURV 45 to calculate interatomic distances
and their root-mean-square variations (σ2). Electron scattering parameters were calculated and

6
Page 7 of 22 Green Chemistry

used to determine the type and number of coordinating atoms, using the Hedin−LundqvistView Article Online
DOI: 10.1039/C9GC03213D

potential.46 Quoted uncertainties on fitted parameters are equal to two standard deviations.

ICP Analysis
Published on 08 November 2019. Downloaded by Centro de Desenvolvimento da Tecnologia Nuclear on 11/11/2019 8:14:34 PM.

The copper and iron concentrations of the samples were determined via ICP analysis using a
Thermo Scientific iCAP Qc ICP-MS. Samples were calibrated against standard solutions
containing known amounts of copper and iron chlorides. Samples were prepared via 1:10 or

Green Chemistry Accepted Manuscript


1:100 dilution in deionised water, after which 2 drops of HNO3 were added to prevent any
precipitation. Where the copper concentration was too low to obtain a good signal-to-noise
ratio in the EXAFS (< 2-5 mM), UV-vis spectra were recorded and compared to spectra of
known species to obtain a good indication of the species present.

Results and discussion


Paint Casting of Cu-based Sulfide Minerals
Paint casting of minerals using DESs was first introduced by Abbott et al.33 as a method to
study the electrochemistry of galena (PbS) without the inclusion of additional binding agents.
It was demonstrated that, even though the mineral itself is relatively insoluble in the DES, the
electrochemical properties of both Pb and S could be studied, since well-defined oxidation and
reduction peaks for both species could be observed. Application of this method to Fe and As
sulfide minerals, such as pyrite and arsenopyrite, showed that the method was applicable to a
range of minerals.34

In the present study, the paint casting method was applied to three Cu-based sulfides and copper
chloride compounds. Figure 2 shows the cyclic voltammograms (CVs) obtained from synthetic
copper sulfides, copper chloride and sodium sulfide. Paint casting CuCl2 produces a CV that is
identical to that of a solution of CuCl2 in 1ChCl:2EG measured at a conventional Pt disc
electrode,47 with average onset potentials of 0.436 V and -0.335 V vs. AgCl reference for the
CuII/I and CuI/0 couples, respectively, highlighting that this method produces results that can be
directly be compared to traditional solution based on CV measurements. Sodium sulfide
displays a redox process at ca. 0.25 V that could potentially be oxidation of S2− to elemental
sulfur or sulfate.

Interestingly, the chalcocite and covellite paint castings show three separate redox couples.
These can be related to the electrochemistry of both Cu and S, in a similar manner to that
7
Green Chemistry Page 8 of 22

obtained from paint casting PbS, where both Pb and S were electrochemically active. 33View
The Article Online
DOI: 10.1039/C9GC03213D

redox couple unrelated to the copper processes is within the same potential range as sulfide
oxidation. Similar peaks were observed during measurements involving pyrite,34 but not the
other iron sulfide minerals such as marcasite or pyrrhotite, or those containing arsenic. This is
Published on 08 November 2019. Downloaded by Centro de Desenvolvimento da Tecnologia Nuclear on 11/11/2019 8:14:34 PM.

important, as it may indicate if oxidation of the sulfide moiety of the mineral is key to the
dissolution mechanism.

Green Chemistry Accepted Manuscript


The corresponding paint casting experiments were also carried out in 1ChCl:2U and 1ChCl:
Ox. However, in 1ChCl:2U the high viscosity (750 cP)48 of the solution resulted in CVs that
were very broad and the individual redox couples could not easily be resolved. During
measurements of covellite in the acidic DES 1ChCl:Ox, a strong H2S-like odour was detected,
and further experiments in the neat liquid were not attempted.

Effect of paste loading and presence of inactive material


In previous work, the mineral galena (PbS) was shown to demonstrate a clear electrochemical
response in the solid state and a linear correlation was observed between the total anodic charge
and the mass of mineral on the anode surface.33 Here, we investigate the effects of paste loading
of the anode, using synthetic covellite in 1ChCl:2EG. Figure 3a shows the electrochemical
response when different masses of covellite are loaded onto the electrode surface. The CVs
retain the same form and potentials as the mass of covellite increases. The total current
increases linearly with covellite loading, indicating that current efficiency remains similar
when using covellite loadings of up to 2 mg on the electrode. However, this data is not
representative of a true ore sample since it does not contain gangue material. Often these
gangue minerals are chemically and electrochemically inert, e.g. silicates or carbonates, and
are not homogeneously distributed throughout ore bodies.

To demonstrate the effect of mineral dilution on the electrochemical response a ‘synthetic’


gangue (alumina) was added to covellite in different proportions, but each experiment had a
total covellite + alumina content of 2 mg. The effect of loading is demonstrated in Figure 3b,
where it can be seen that the overall shape of the CVs and values of electrode potentials are not
significantly shifted by the presence of this gangue material. CVs of covellite and chalcocite
(Figure 4) were almost identical to those of the synthetic material, indicating that
electrochemically inactive gangue materials do not adversely affect the electrochemical
behaviour of the copper sulfide other than by dilution or increasing the resistivity of the paste.
8
Page 9 of 22 Green Chemistry

View Article Online


DOI: 10.1039/C9GC03213D

Electrochemical dissolution of chalcopyrite (CuFeS2)


In aqueous systems chalcopyrite is significantly more difficult to dissolve chemically than
covellite and chalcocite due to surface effects, such as the generation of polysulfide species,9,
Published on 08 November 2019. Downloaded by Centro de Desenvolvimento da Tecnologia Nuclear on 11/11/2019 8:14:34 PM.

49 with leaching processes commonly requiring high temperatures or high pressures.5, 50 The
presence of Cl− ions has been suggested to alter the morphology of any surface sulfur formed,
allowing the chalcopyrite to leach more easily by preventing passivation.51 Additionally, the

Green Chemistry Accepted Manuscript


influence of iron species must be considered, as the formation of insoluble iron
oxide/hydroxide phases in aqueous sulfate systems has been observed if the system was heated
or electrolysed too fast.52

DESs provide a high Cl‒ ion content and low water content, whilst reducing the need for either
high additional salt or acid concentrations, circumventing most oxide chemistry.30 Therefore,
they should provide a suitable environment for chalcopyrite dissolution. Paint casting was
carried out directly on the natural material, and the CV of chalcopyrite resembles that of
covellite more than chalcocite, displaying the oxidation and reduction peaks associated with
CuII/I (0.422 V), along with the reduction peak for CuI/0 (-0.354 V, onset only) but not the
corresponding stripping peak (Figure 4c). In aqueous leaching systems, it has been suggested
that chalcopyrite is first converted to Cu2S, and additional peaks in the CV are related to a
mixture of other Cu-sulfide species, such as Cu1.92S, Cu1.60S, and CuS.9 Alternatively, the
presence of chloride in aqueous leachates is known to change surface sulfur into a more
crystalline and porous form,51 which may also take place in DES media. The lack of defined
peaks for the FeIII/II couple is most likely due to an overlap with the CuII/I redox couple as both
redox couples have similar electrode potentials (see Figure S2, ESI).

Metal speciation after electrochemical dissolution in DESs


To determine the speciation of metal ions after electrochemical dissolution of the Cu-sulfide
minerals, cyclic voltammetry (CV), along with a combination of UV-Vis and extended X-ray
absorption fine structure (EXAFS) spectroscopies were employed. All three minerals were
anodically dissolved in 1ChCl:2EG for a total of 24 h at 1.5 V, then the samples were filtered
to remove any particulate matter. Bulk dissolution of both covellite and chalcocite produce
very strongly coloured yellow solutions (Figure 5, inset), characteristic of the [CuCl4]2˗
complex.53

9
Green Chemistry Page 10 of 22

After bulk dissolution of chalcopyrite, a light green/yellow solution was produced.


DOI: From the
View Article Online
10.1039/C9GC03213D

UV-vis spectra, different species are expected to be present. Due to the pale colour of the
solution and change in wavelength of the absorption maxima, the presence of a CuI species
such as [CuCl2]− was considered.53 As ICP analysis showed a 1:1 ratio of Cu:Fe in solution,
Published on 08 November 2019. Downloaded by Centro de Desenvolvimento da Tecnologia Nuclear on 11/11/2019 8:14:34 PM.

solutions were made containing a 1:1 mixture of CuCl with either FeCl3 or FeCl2. The spectrum
of the CuCl: FeCl2 mix matched the closest, indicating that iron was most likely in the +2
oxidation state, suggesting that FeIII has been reduced, either at the cathode, or through the

Green Chemistry Accepted Manuscript


oxidation of sulfide.

As UV-vis could only provide indicative information about speciation, EXAFS was used to
determine the speciation of copper ions produced via electrodissolution of three copper
sulfides. From X-ray absorption near edge spectroscopy (XANES), it is possible to obtain an
indication of which oxidation states are present.

EXAFS study of solubilised copper sulfide based minerals


Previous investigations into the speciation of high purity metal chloride salts in IL and DES
media revealed that coordination was, in general, controlled by the anionic component of the
liquid.53 However, it is possible that metal speciation after electrolytic dissolution of natural
minerals may be altered due to the presence of decomposition products from the solvent or the
other components in the minerals, such as anionic sulfur species. Multinuclear species may
also be formed under these conditions.

For the solutions formed following anodic dissolution of covellite and chalcocite in
1ChCl:2EG, tetrachloride species with Cu-Cl path lengths of 2.24 – 2.25 Å were present,
similar to solutions of CuCl2 in 1ChCl:2EG (Table 1).47, 53 XANES data (Figure 6) confirms
that CuII is present in both cases, i.e. the CuI in the minerals has been oxidised. Due to the
sulfide content of the samples, it is likely that various sulfur-containing ligands will be present
in solution. EXAFS cannot easily distinguish between Cl- and S-coordination; however,
formation of the Cl-species is most likely due to the high ca. 4.2 M chloride content of the
DES.

Contrary to the two copper sulfide systems, a CuI species was indicated for the anodic
dissolution of chalcopyrite. Coordination number, Cu-Cl path lengths, and thermal disorder
were all of a similar magnitude to the parameters obtained from a CuCl solution in the same
10
Page 11 of 22 Green Chemistry

DES. Due to the Fe K-edge being softer, a much higher iron concentration is required to obtain
View Article Online
DOI: 10.1039/C9GC03213D

data with a good signal-to-noise ratio, therefore we cannot confirm the speciation or oxidation
state of iron in solution using EXAFS. However, it is known that iron chloride salts form either
[FeCl4]2− or [FeCl4]−, depending on the iron oxidation state (Table 2). As the method of
Published on 08 November 2019. Downloaded by Centro de Desenvolvimento da Tecnologia Nuclear on 11/11/2019 8:14:34 PM.

dissolution has no apparent effect on the coordination of copper ions in 1ChCl:2EG, we


propose that any soluble iron species formed during electrodissolution are also likely to be
chloride coordinated. However, the possibility of Fe-S or Fe-O complexes cannot be ruled out.

Green Chemistry Accepted Manuscript


As no Cu-Fe paths were observed in any of the Cu-edge data, it is unlikely that mixed-metal
species are present in solution.

Near-edge measurements indicated that the oxidation state of the electrodissolved copper
changed based on the other species present in the mineral. We propose that this change in
copper oxidation state of the solution species can be explained via electrode potentials. In
1ChCl:2EG, the formal electrode potentials for the CuII/I and FeIII/II couples are +0.349 V and
+0.293 V vs a [Fe(CN)6]3−/4− internal standard, respectively (see Figure S2, ESI), i.e. the FeIII
complex is stabilised by approximately 56 mV compared to the CuII complex. It is therefore
possible that any FeII formed at the cathode during electrolysis reduces CuII to CuI.
Alternatively, FeIII may oxidise S2‒ to S0 or SO42‒, which may in turn interact with the CuI ions,
as well as forming further FeII that may reduce CuII. No iron is present in either the CuS or
Cu2S systems, hence the copper species retain the higher oxidation state.

For copper chloride salts, speciation in 1ChCl:2U and 1ChCl:2EG is similar. However, after
anodic dissolution of covellite in 1ChCl:2U, a mixed ligand complex formed, containing 1.4(3)
x O-/N-donor and 1.9(4) x Cl‒ ligands. This O-/N-donor ligand is likely to be based on urea,
or decomposition products such as ammonia. The absence of multiple scattering signals around
4-5 Å would tend to suggest that urea is not directly coordinated to the copper ions, as was seen
for CrCl3·6H2O:urea Type 4 eutectics.54 Spectra for chalcocite were not recorded due to low
Cu concentrations. From comparison of UV-vis spectra, we propose that the speciation is the
same as for covellite. Intriguingly, linear combination fitting of the XANES data suggests that
copper ions are present in a ratio of 30% CuI and 70% CuII (see Figure S3, ESI).

Table 2 shows that when urea was present in the DES, FeCl2 and FeCl3 form O- or N-donor
complexes. In these cases, the coordinating ligand is probably urea, due to multiple scattering
signals around 4-5 Å. Interestingly, the fitted parameters and XANES data indicates that FeIII-
11
Green Chemistry Page 12 of 22

species are present in both cases (see Figure S4, ESI). If Fe species formed from the
DOI: anodic
View Article Online
10.1039/C9GC03213D

dissolution of chalcopyrite behave in the same way, the use of 1ChCl:2U should result in
different speciation than 1ChCl:2EG and hence electrochemistry. As discussed previously,
1ChCl:2U was too viscous to be a viable solvent for bulk electrodissolution of chalcopyrite.
Published on 08 November 2019. Downloaded by Centro de Desenvolvimento da Tecnologia Nuclear on 11/11/2019 8:14:34 PM.

Therefore, in order to retain these altered speciation properties whilst reducing viscosity and
increasing conductivity, a solution of 1ChCl:2EG with 20 wt% 1ChCl:2U was used.

Green Chemistry Accepted Manuscript


As EXAFS for the 1ChCl:Ox systems was not measured, the UV-vis spectra were once again
considered. By comparing the unknown 1ChCl:Ox systems to the known 1ChCl:2EG and
1ChCl:2U systems, it can be seen that copper retains the tetrachloride speciation, whilst iron
forms a completely different complex in each solvent (Figure S5, ESI). Although the
speciation behaviours of metal salts in DESs are not directly comparable to aqueous solutions,
aqueous stability constants can still be used to obtain an indication of which complex is most
likely to be dominant in DES media. In aqueous solutions, the cumulative stability constants
for the tetrachloride and trisoxalato complexes at 25°C are -1.26 and 18.49, respectively.55, 56
The stability constant for the tetrachloride species in the DES systems will naturally be higher
than the aqueous value, given the EXAFS data for iron chloride salts in 1ChCl:2EG, however
the difference between the stability constants is sufficiently large that it is most likely that iron
oxalate complexes dominate in 1ChCl:Ox.

Bulk recovery of copper from chalcopyrite


A mixed DES system of 1ChCl:1.5EG:0.5U DESs has already been applied to the treatment of
electric arc furnace dust, selectively recovering lead and zinc from a dust containing Pb, Zn
and Cd.36 By mixing DESs, a solvent with optimal viscosity and metal solubility properties can
be obtained. In the present study, three DESs were tested for bulk leaching and recovery
experiments; neat 1ChCl:2EG, 1ChCl:2EG with 20 wt% 1ChCl:2U and 1ChCl:2EG with 20
wt% 1ChCl:1Ox. To check that no H2S was generated during the experiment using the 20 wt%
1ChCl:1Ox system, a strip of lead acetate paper was suspended above the experiment. No
blackening of the paper was observed, indicating that H2S is not generated from this mixture.

For all systems, the Ni cathode in the cell (Figure 1) was replaced every 24 h and the
composition of the material deposited on the cathode was determined. Figure 7 shows that
with the 1ChCl:2EG system, the first electrowon deposit obtained after 24 h was almost pure
copper. This showed that the recovery process is selective. However, with subsequent digestion
12
Page 13 of 22 Green Chemistry

and plating, the iron content gradually increased while that for copper decreased. The
View Article Online
DOI: 10.1039/C9GC03213D

increasing iron content was coupled with a corresponding increase of the oxygen content (but
negligible sulfur), indicating the formation of iron oxides. The addition of 20 wt% 1ChCl:2U
improved the selective recovery of copper, with a maximum of 3.4 at% Fe impurity on one
Published on 08 November 2019. Downloaded by Centro de Desenvolvimento da Tecnologia Nuclear on 11/11/2019 8:14:34 PM.

plate. The best results were obtained from the 20 wt% 1ChCl:1Ox system, where only 1.7 at%
Fe was present on the final plate, with Cu averaging 99% over the five plates. By altering the
HBDs present, the speciation (and hence complex stability) of the Fe complexes are changed,

Green Chemistry Accepted Manuscript


without affecting Cu deposition. This in turn alters the reduction properties of the Fe species
in solution.

The water content of the solvent can affect speciation, however according to De Vreese et al.57
and Li et al.58 the amount of water necessary to cause a significant change in copper speciation
in DES or ILs is in the region of ca. 40 wt% H2O. As we only anticipate a maximum of ca. 5-
10 wt% H2O even from very long exposure to atmospheric moisture or damp minerals, the only
impacts expected would be a decrease in solution viscosity, and corresponding increase in mass
transport properties.59 This lack of water sensitivity is important for industrially relevant
processes, where the need for rigidly controlled atmospheres are undesirable.

After the neat 1ChCl:2EG experiment, the electrolyte surrounding the cathode turned brown,
with a darker brown precipitate at the bottom of the container. The solutions containing 20 wt%
1ChCl:2U and 1ChCl:1Ox instead turned green and grey/blue respectively, with a green/brown
bottom precipitate. The high sulfur content in the 1ChCl:2EG bottom precipitate (Table 3)
compared to the other systems suggests that either elemental sulfur or iron sulfides are present
in the deposit. XRD analysis indicated that the precipitates were composed of amorphous
material, with some unreacted chalcopyrite present in the 20 wt% 1ChCl:1Ox solution.
Addition of aqueous BaCl2 solution to the remaining liquid produced a cloudy precipitate for
all of the DESs used, indicating that some of the sulfur forms sulfate, as it would in an aqueous
solution. As either urea or oxalic acid is added to 1ChCl:2EG, the total mass of bottom
precipitate decreases. The mass of copper remains roughly constant but that of iron decreases
significantly. This shows that additives can significantly change solubility and selectivity for
extraction.

Subsequent to the completion of these experiments, it was discovered that carboxylic acid-
based DESs degrade via an esterification mechanism between the acidic HBD and the OH-
13
Green Chemistry Page 14 of 22

group of the choline chloride.60 This esterification will naturally change the properties ofViewthe
Article Online
DOI: 10.1039/C9GC03213D

employed DES over time and will be of significance for the reuse and recyclability of these
solvents.
Published on 08 November 2019. Downloaded by Centro de Desenvolvimento da Tecnologia Nuclear on 11/11/2019 8:14:34 PM.

Conclusions
The electrodissolution of chalcocite, covellite and chalcopyrite was investigated in DES media.
All experiments resulted in the release of copper ions which could be recovered from solution.

Green Chemistry Accepted Manuscript


No surface passivation or formation of elemental sulfur was observed during these
experiments. The main copper species present in 1ChCl:2EG after dissolution of chalcocite
and covellite was [CuCl4]2−. In the solution formed from chalcopyrite, a mix of CuII and CuI
species were formed instead, potentially due to generation of reducing FeII species at the
cathode. In 1ChCl:2U, copper had a mixed chloride/O- or N-donor coordination, potentially
altering electrochemical behaviour.

Electrochemical oxidation of sulfide minerals resulted in the formation of sulfate in solution


and avoided the generation of SO2 or H2S, except when 1ChCl:1Ox was used as a neat DES.
Copper could be electrowon from chalcopyrite in the three DESs tested with varying purities.
In 1ChCl:2EG the proportion of iron increased over the course of the experiment. With the
addition of 20 wt% 1ChCl:2U, copper purity increased, and the best selective recovery of
copper (99 at%) was with 20 wt% 1ChCl:1Ox solution. This is inferred to be due to changes
in iron speciation and electrochemical behaviour between the three liquids, resulting in
different electrodeposition conditions being required for Fe. From analysis of the liquids left
after the experiment and their precipitates, it can be shown that electrolytic dissolution alone
does not selectively separate the two metals, and that selectivity is instead obtained at the
cathode during electrowinning. This demonstrates that DESs can be used to selectively recover
metals from solution by changing solute speciation and hence redox properties.

Acknowledgements
The authors would like to thank the German Research Foundation (DFG grant FR 3458/2-1)
and the NERC Minerals Security of Supply (SoS) grant NE/M010848/1 Tellurium and
Selenium Cycling and Supply (TeaSe) for funding, and the European Synchrotron Radiation
Facility for beamtime. The authors would also like to thank Dr Philip Bird for carrying out the
Zeiss analysis and Vicky Ward, the Geological Curator at Leicester, for helping source the
samples.
14
Page 15 of 22 Green Chemistry

View Article Online


DOI: 10.1039/C9GC03213D

This paper is dedicated to Syahrie Angarra for his hard work during his Masters project who
sadly passed away after completing his Masters dissertation. He will always be remembered as
a warm-hearted colleague.
Published on 08 November 2019. Downloaded by Centro de Desenvolvimento da Tecnologia Nuclear on 11/11/2019 8:14:34 PM.

Author information
ORC-IDs

Green Chemistry Accepted Manuscript


Andrew P. Abbott: 0000-0001-9556-8341
Gero Frisch: 0000-0002-8997-7744
Robert Harris: 0000-0002-2537-8240
Jennifer Hartley: 0000-0003-2174-4458
Gawen R. T. Jenkin: 0000-0001-9202-7128

15
Green Chemistry Page 16 of 22

Figures and Tables View Article Online


DOI: 10.1039/C9GC03213D
Published on 08 November 2019. Downloaded by Centro de Desenvolvimento da Tecnologia Nuclear on 11/11/2019 8:14:34 PM.

Green Chemistry Accepted Manuscript


Figure 1: The cell design used in the experiment. Left: Before the experiment and right:
after the experiment was completed showing the different layers of material present.
Colours differed depending on DES used.

1.5 a3 1.5 0.6

a3
1.0 1.0 0.4
a2 a2
a1
0.5 0.5 0.2
i / mA

i / mA

i / A

0.0 0.0 a1 0.0

-0.5 -0.5 c3 -0.2


c3 c1 c1
c2 c2
-1.0 -1.0 -0.4
-0.8 -0.4 0.0 0.4 0.8 -0.8 -0.4 0.0 0.4 0.8
E / V vs Ag/AgCl E / V vs Ag/AgCl

a) Chalcocite b) Covellite
Figure 2: CVs of (a) synthetic chalcocite (Cu2S) with overlaid CuCl2 (blue), and (b)
synthetic covellite (CuS) with overlaid Na2S (red) obtained via paint casting onto a Pt flag
electrode in 1ChCl:2EG, with Ag/AgCl reference electrode.

16
Page 17 of 22 Green Chemistry

View Article Online


CuS : alumina DOI: 10.1039/C9GC03213D
2.0 1.5
0.4 mg 1:4
1.5 0.8 mg 2:3
1.2 mg 1.0 3:2
1.0 2.0 mg 4:1

0.5 0.5
i / mA

i / mA
Published on 08 November 2019. Downloaded by Centro de Desenvolvimento da Tecnologia Nuclear on 11/11/2019 8:14:34 PM.

0.0
0.0
-0.5
-0.5
-1.0

-1.5

Green Chemistry Accepted Manuscript


-1.0
-0.8 -0.4 0.0 0.4 0.8 -0.8 -0.4 0.0 0.4 0.8
E / V vs Ag/AgCl E / V vs Ag/AgCl

Figure 3: Voltammograms of covellite at different loadings (left), and effect of covellite to


alumina loading (right). All measurements carried out in 1ChCl:2EG.

1.5 1.5 0.20


a3
a3
a3 0.15
1.0 1.0 a2

a2 0.10
0.5 0.5
a2
i / mA

i / mA

i / mA
0.05
a1 a1
0.0 0.0 a1
0.00

c3
-0.5 -0.5 c3 -0.05
c1 c1
c2 c2 c1
-0.10 c3
c2
-1.0 -1.0
-0.8 -0.4 0.0 0.4 0.8 -0.8 -0.4 0.0 0.4 0.8 -0.8 -0.4 0.0 0.4 0.8
E / V vs Ag/AgCl E / V vs Ag/AgCl E / V vs Ag/AgCl

Figure 4: Voltammograms of a) natural covellite, b) natural chalcocite, and c) natural


chalcopyrite in 1ChCl:2EG. All were pasted onto the anode surface (Pt flag electrode).

Cu2S CuS CuFeS2


Absorbance

200 300 400 500 600 700


 / nm

Figure 5: UV-Vis Spectroscopy of solutions generated from bulk anodic dissolution of


covellite (CuS black), chalcocite (Cu2S red), and chalcopyrite (CuFeS2 blue) in
1ChCl:2EG.

17
Green Chemistry Page 18 of 22

View Article Online


DOI: 10.1039/C9GC03213D

CuCl2
Published on 08 November 2019. Downloaded by Centro de Desenvolvimento da Tecnologia Nuclear on 11/11/2019 8:14:34 PM.

CuS
Cu2S
normalised x(E)

FT k2 (k)

Green Chemistry Accepted Manuscript


8970 8980 8990 9000 9010 9020 9030 0 1 2 3 4 5 6
Energy (eV) r/Å

CuCl
CuFeS2
normalised x(E)

FT k (k)
2

8970 8980 8990 9000 9010 9020 9030 0 1 2 3 4 5 6


Energy (eV) r/Å

Figure 6: XANES region (left) and k2-weighted Fourier transform (right) for CuCl2,
covellite (CuS), and chalcocite (Cu2S) (top), and for CuCl and chalcopyrite (CuFeS2)
(bottom) in 1ChCl:2EG. Data are circles and fits are lines. Spectra are offset for clarity.

Fe
Cu
100 100 100
at % composition

at % composition

at % composition

50 50 50

0 0 0
1 2 3 4 5 1 2 3 4 5 1 2 3 4 5
Plate number Plate number Plate number

Figure 7: The composition of the deposit on five consecutive cathode plates from each
experiment for direct electrowinning from chalcopyrite. Left: 1ChCl:2EG only, Middle: +
20 wt% 1ChCl:2U, and Right: + 20 wt% 1ChCl:1Ox.

18
Page 19 of 22 Green Chemistry

Table 1: EXAFS fit parameters for solutions of copper salts and mineral dissolution inView Article Online
DOI: 10.1039/C9GC03213D

1ChCl:2EG (top) and 1ChCl:2U (bottom).


Solute Coordinating Number of Distance from Debye-Waller Fit index,
atom/group atoms, N Cu, r/Å factor, a (Å2) R1/%
Published on 08 November 2019. Downloaded by Centro de Desenvolvimento da Tecnologia Nuclear on 11/11/2019 8:14:34 PM.

CuCl2 * Cl 3.7(2) 2.252(5) 0.008(1) 4.48 %


CuCl * Cl 2.4(2) 2.190(9) 0.016(3) 8.08 %
CuS (covellite) Cl 4.4(2) 2.248(4) 0.012(1) 2.74 %

Green Chemistry Accepted Manuscript


Cu2S (chalcocite) Cl 4.3(4) 2.239(7) 0.010(2) 5.87 %
CuFeS2 (chalcopyrite)§ Cl 3.0(4) 2.21(1) 0.019(5) 15.21 %
Solute Coordinating Number of Distance from Debye-Waller Fit index,
atom/group atoms, N Cu, r/Å factor, a (Å2) R1/%
CuCl2 * Cl 3.9(3) 2.239(7) 0.013(2) 6.39 %
CuCl * Cl 2.5(2) 2.224(7) 0.012(2) 7.55 %
CuS (covellite) O 1.4(3) 1.95(2) 0.007(3) 1.42 %
Cl 1.9(4) 2.24(1) 0.015(3)
* Data already presented in reference 53
§ Data from single scan

Table 2: EXAFS fit parameters for solutions of iron salts in 1ChCl:2EG (top) and
1ChCl:2U (bottom).
Solute Coordinating Number of Distance from Debye-Waller Fit index,
atom/group atoms, N Fe, r/Å factor, a (Å2) R1/%
FeCl2 * Cl 3.8(2) 2.313(3) 0.0065(8) 2.88 %
FeCl3 Cl 3.8(1) 2.205(3) 0.0042(7) 1.58 %
Solute Coordinating Number of Distance from Debye-Waller Fit index,
atom/group atoms, N Fe, r/Å factor, a (Å2) R1/%
FeCl2 * O 4.8(4) 2.08(1) 0.019(3) 6.84 %
FeCl3 O 4.8(3) 2.066(9) 0.017(3) 6.06 %
* Data already presented in reference 53

Table 3: Composition of the precipitate remaining in cell after electrowinning.

System Cu / g Fe / g S/g
1ChCl:2EG 0.22 0.53 0.76
+ 20 wt% 1ChCl:2U 0.45 0.19 0.04
+ 20 wt% 1ChCl:1Ox 0.33 0.033 0.06

19
Green Chemistry Page 20 of 22

References View Article Online


DOI: 10.1039/C9GC03213D

1. H. R. Watling, Hydromet., 2013, 140, 163–180.


2. Y. Li, N. Kawashima, J. Li, A. P. Chandra and A. R. Gerson, Adv. Colloid Interface Sci., 2013,
197-198, 1-32.
3. ICSG, The World Copper Factbook 2018.
Published on 08 November 2019. Downloaded by Centro de Desenvolvimento da Tecnologia Nuclear on 11/11/2019 8:14:34 PM.

4. F. Habashi, Hydromet., 2005, 79, 15-22.


5. F. Habashi, Min. Proc. Extract. Metal. Rev., 2007, 15, 5-12.
6. Y. S. R. Hara, Min. Proc. Extract. Metal. Rev., 2015, 36, 26-38.
7. E. M. Córdoba, J. A. Muñoz, M. L. Blázquez, F. González and A. Ballester, Hydromet., 2008,
93, 81-87.

Green Chemistry Accepted Manuscript


8. M. Lundström, J. Aromaa, O. Forsén and M. H. Barker, Can. Met. Quart., 2008, 47, 245-252.
9. H.-b. Zhao, M.-h. Hu, Y.-n. Li, S. Zhu, W.-q. Qin, G.-z. Qiu and J. Wang, Trans. Nonfer. Metals
Soc. China, 2015, 25, 303−313.
10. R. G. McDonald and D. M. Muir, Hydromet., 2007, 86, 191-205.
11. D. G. Dixon, D. D. Mayne and K. G. Baxter, Can. Met. Quart., 2008, 47, 327-336.
12. N. Hiroyoshi, H. Miki, T. Hirajima and M. Tsunekawa, Hydromet., 2001, 60, 185–197.
13. M. M. Antonijević, Z. D. Janković and M. D. Dimitrijević, Hydromet., 2004, 71, 329–334.
14. H. He, J.-L. Xia, Y. Yang, H. Jiang, C.-q. Xiao, L. Zheng, C.-y. Ma, Y.-d. Zhao and G.-z. Qiu,
Hydromet., 2009, 99, 45-50.
15. H. Niedermeyer, J. P. Hallett, I. J. Villar-Garcia, P. A. Hunt and T. Welton, Chem Soc Rev,
2012, 41, 7780-7802.
16. K. Yoo, S.-k. Kim, J.-c. Lee, M. Ito, M. Tsunekawa and N. Hiroyoshi, Min. Eng., 2010, 23,
471-477.
17. O. Kuzmina, E. Symianakis, D. Godfrey, T. Albrecht and T. Welton, Phys. Chem. Chem. Phys.,
2017, 19, 21556-21564.
18. J. A. Whitehead, J. Zhang, N. Pereira, A. McCluskey and G. A. Lawrance, Hydromet., 2007,
88, 109-120.
19. C. L. Aguirre, N. Toro, N. Carvajal, H. Watling and C. Aguirre, Min. Eng., 2016, 99, 60-66.
20. T. Dong, Y. Hua, Q. Zhang and D. Zhou, Hydromet., 2009, 99, 33-38.
21. J. A. Whitehead, G. A. Lawrance and A. McCluskey, Green Chem., 2004, 6, 313-315.
22. C. Carlesi, E. Cortes, G. Dibernardi, J. Morales and E. Muñoz, Hydromet., 2016, 161, 29-33.
23. A. Al-Zubeidi, D. Godfrey and T. Albrecht, J. Electroanal. Chem., 2018, 819, 130-135.
24. J. Hu, G. Tian, F. Zi and X. Hu, Hydromet., 2017, 169, 1-8.
25. R. Winand, Hydromet., 1991, 27, 285-316.
26. J. E. Dutrizac, Hydromet., 1992, 29, 1-45.
27. Z. Y. Lu, M. I. Jeffrey and F. Lawson, Hydromet., 2000, 56, 145-155.
28. V. J. Martínez-Gómez, J. C. Fuentes-Aceituno, R. Pérez-Garibay and J.-c. Lee, Hydromet.,
2018, 181, 195–205.
29. J. Liddicoat and D. Dreisinger, Hydromet., 2007, 89, 323-331.
30. A. P. Abbott, G. Frisch, J. Hartley and K. S. Ryder, Green Chem., 2011, 13, 471-481.
31. G. R. T. Jenkin, A. Z. M. Al-Bassam, R. C. Harris, A. P. Abbott, D. J. Smith, D. A. Holwell,
R. J. Chapman and C. J. Stanley, Min. Eng., 2016, 87, 18-24.
32. D. Jones, J. Hartley, G. Frisch, M. Purnell and L. Darras, Palaeontologia Electronica, 2012,
15.2.4T, 1-7.
33. A. P. Abbott, F. Bevan, M. Baeuerle, R. C. Harris and G. R. T. Jenkin, Electrochem. Comm.,
2017, 76, 20-23.
34. A. P. Abbott, A. Z. M. Al-Bassam, A. Goddard, R. C. Harris, G. R. T. Jenkin, F. J. Nisbet and
M. Wieland, Green Chem., 2017, 19, 2225-2233.
35. A. Entezari-Zarandi and F. Larachi, J. Rare Earths, 2019, 37, 528-533.
36. A. P. Abbott, J. Collins, I. Dalrymple, R. C. Harris, R. Mistry, F. Qiu, J. Scheirer and W. R.
Wise, Aust. J. Chem., 2009, 62, 341–347.
37. P. Zürner and G. Frisch, ACS Sus. Chem. Eng., 2019, 7, 5300-5308.
38. C. I. Pearce, R. A. D. Pattrick, D. J. Vaughan, C. M. B. Henderson and G. van der Laan,
Geochim. Cosmochim. Acta, 2006, 70, 4635-4642.

20
Page 21 of 22 Green Chemistry

39. S. W. Goh, A. N. Buckley and R. N. Lamb, Minerals Eng., 2006, 19, 204-208. DOI: 10.1039/C9GC03213D
View Article Online

40. H. T. Evans and J. A. Konnert, Am. Mineral., 1976, 61, 996-1000.


41. A. P. Abbott, D. Boothby, G. Capper, D. L. Davies and R. K. Rasheed, J. Am. Chem. Soc.,
2004, 126, 9142-9147.
42. A. P. Abbott, G. Capper, D. L. Davies, R. K. Rasheed and V. Tambyrajah, Chem. Commun.,
2003, 38, 70-71.
Published on 08 November 2019. Downloaded by Centro de Desenvolvimento da Tecnologia Nuclear on 11/11/2019 8:14:34 PM.

43. S. Nikitenko, A. M. Beale, A. M. van der Eerden, S. D. Jacques, O. Leynaud, M. G. O'Brien,


D. Detollenaere, R. Kaptein, B. M. Weckhuysen and W. Bras, J. Synchrotron Rad., 2008, 15,
632-640.
44. B. Ravel and M. Newville, J. Synchrotron Rad., 2005, 12, 537-541.
45. S. Tomić, B. G. Searle, A. Wander, N. M. Harrison, A. J. Dent, J. F. W. Mosselmans and J. E.

Green Chemistry Accepted Manuscript


Inglesfield, CCLRC Technical Report, 2004, DL-TR-2005-001.
46. S. J. Gurman, N. Binsted and I. Ross, J. Phys. Chem. C, Solid State Phys., 1986, 19, 1845-1861.
47. A. P. Abbott, G. Frisch, S. J. Gurman, A. R. Hillman, J. Hartley, F. Holyoak and K. S. Ryder,
Chem. Commun., 2011, 47, 10031-10033.
48. C. D'Agostino, R. C. Harris, A. P. Abbott, L. F. Gladden and M. D. Mantle, Phys. Chem. Chem.
Phys., 2011, 13, 21383-21391.
49. A. Ghahremaninezhad, D. G. Dixon and E. Asselin, Electrochim. Acta, 2013, 87, 97– 112.
50. D. Dreisinger and N. Abed, Hydromet., 2002, 66, 37–57.
51. L. Pikna, L. Lux and T. Grygar, Chem. Pap., 2006, 60, 293-296.
52. E. M. Córdoba, J. A. Muñoz, M. L. Blázquez, F. González and A. Ballester, Min. Eng., 2009,
22, 229-235.
53. J. M. Hartley, C. M. Ip, G. C. Forrest, K. Singh, S. J. Gurman, K. S. Ryder, A. P. Abbott and
G. Frisch, Inorg. Chem., 2014, 53, 6280-6288.
54. A. P. Abbott, A. A. Al-Barzinjy, P. D. Abbott, G. Frisch, R. C. Harris, J. Hartley and K. S.
Ryder, Phys. Chem. Chem. Phys., 2014, 16, 9047-9055.
55. W. Liu, B. Etschmann, J. Brugger, L. Spiccia, G. Foran, B. McInnes, Chem. Geol., 2006, 231,
326-349.
56. A. E. Martell, R. M. Smith, Critical stability constants volume 3: Other Organic Ligands,
Plenum Press, New York (1977).
57. P. De Vreese, N. R. Brooks, K. Van Hecke, L. Van Meervelt, E. Matthijs, K. Binnemans and
R. Van Deun, Inorg. Chem., 2012, 51, 4972-4981.
58. G. Li, D. M. Camaioni, J. E. Amonette, Z. C. Zhang, T. J. Johnson and J. L. Fulton, J. Phys,
Chem. B, 2010, 114, 12614-12622.
59. A. Y. M. Al-Murshedi, J. M. Hartley, A. P. Abbott and K. S. Ryder, Trans. Inst. Metal Finish.,
2019 (in print).
60. N. Rodriguez Rodriguez, A. van den Bruinhorst, L. J. B. M. Kollau, M. C. Kroon and K.
Binnemans, ACS Sus. Chem. Eng., 2019, 7, 11521-11528.

21
Green Chemistry Page 22 of 22

Direct Extraction of Copper from Copper Sulfide Minerals using


DOI:Deep
View Article Online
10.1039/C9GC03213D

Eutectic Solvents
Syahrie Anggara,a Francesca Bevan,a Robert C. Harris,a Jennifer M. Hartley,a,b Gero Frisch,b
Published on 08 November 2019. Downloaded by Centro de Desenvolvimento da Tecnologia Nuclear on 11/11/2019 8:14:34 PM.

Gawen R. T. Jenkin,c Andrew P. Abbotta*

a School of Chemistry, University of Leicester, Leicester, LE1 7RH, UK

Green Chemistry Accepted Manuscript


b Institut für Anorganische Chemie, TU Bergakademie Freiberg, 09599 Freiberg, Germany
c School of Geography, Geology and the Environment, University of Leicester, Leicester,
LE1 7RH, UK

Graphical Abstract
Copper can be extracted from the main copper mineral, chalcopyrite, without the formation
of hydrogen sulfide or sulfur dioxide

Das könnte Ihnen auch gefallen