New insights into the influence of redox potential on chalcopyrite leaching behaviour

New insights into the influence of redox potential on chalcopyrite leaching behaviour

Minerals Engineering 100 (2017) 9–16 Contents lists available at ScienceDirect Minerals Engineering journal homepage: www.elsevier.com/locate/mineng...

1MB Sizes 0 Downloads 43 Views

Minerals Engineering 100 (2017) 9–16

Contents lists available at ScienceDirect

Minerals Engineering journal homepage: www.elsevier.com/locate/mineng

New insights into the influence of redox potential on chalcopyrite leaching behaviour Mohammad Khoshkhoo a,⇑, Mark Dopson b, Fredrik Engström a, Åke Sandström a a b

MiMeR – Minerals and Metallurgical Research Laboratory, Luleå University of Technology, SE-971 87 Luleå, Sweden Centre for Ecology and Evolution in Microbial Model Systems, Linnaeus University, 391 82 Kalmar, Sweden

a r t i c l e

i n f o

Article history: Received 2 July 2016 Revised 28 September 2016 Accepted 2 October 2016

Keywords: Chalcopyrite Hindered dissolution Passivation Reductive leaching Galvanic interaction

a b s t r a c t Chalcopyrite (CuFeS2) is the most economically important and most refractory copper mineral when treated in conventional sulphate media leaching systems. In this study, the effect of solution redox potential on leaching of a pure and a pyritic chalcopyrite concentrate was investigated using concentrates with fresh and aged surfaces. In experiments using concentrates with fresh surfaces, the response to redox potential depended on the presence of pyrite: fresh pyritic concentrate leached more effectively at low redox potential (in agreement with reductive leaching mechanisms), while the leaching efficiencies from fresh pure concentrate were similar at high and low redox potentials. The data suggested that the reductive leaching mechanism does not necessarily result in higher and faster recoveries in the absence of the galvanic interaction induced by the presence of pyrite. It was also found that exposure of chalcopyrite to atmospheric oxidation prior to leaching (ageing) had an effect on leaching behaviour in response to redox potential: copper recoveries in leaching of aged concentrates were higher at high redox potentials. This behaviour was attributed to the presence of iron–oxyhydroxides on the surface of aged concentrates. Based on the data from this investigation and previous surface studies, it is proposed that iron– oxyhydroxides play an important role in triggering the hindered dissolution of chalcopyrite. Ó 2016 Published by Elsevier Ltd.

1. Introduction As the most abundant copper mineral, chalcopyrite (CuFeS2) is the most economically important copper resource. Leaching of chalcopyrite in sulphate media is mainly preferred for its easy handling and compatibility with established chalcocite (Cu2S) and covellite (CuS) heap leaching/bioleaching processes as well as with conventional solvent extraction and electrowinning technologies (Marsden, 2007). However, the slow dissolution of chalcopyrite at ambient pressure in sulphate media has hindered its industrial application. Nevertheless, due to the continuous depletion of high grade ores, future hydrometallurgical alternatives for chalcopyrite treatment will increase in importance and applications. As a result, different aspects of chalcopyrite leaching have been extensively researched in recent decades. A summary of these studies can be found in Debernardi and Carlesi (2013) and Li et al. (2013). A review of hydrometallurgical options for chalcopyrite treatment is available in Dreisinger (2006) amongst others. The slow dissolution of chalcopyrite is termed ‘passivation’ or ‘hindered dissolution’ and is believed to be due to formation of ⇑ Corresponding author. E-mail address: [email protected] (M. Khoshkhoo). http://dx.doi.org/10.1016/j.mineng.2016.10.003 0892-6875/Ó 2016 Published by Elsevier Ltd.

compounds on the mineral surface during leaching. The most frequently suggested candidates to form the passivating layers are metal-deficient phases, elemental sulphur and jarosite (Klauber, 2008). Recently iron–oxyhydroxides have also been proposed to cause passivation (Khoshkhoo et al., 2014a). Similarly, other aspects of chalcopyrite (bio)leaching, such as the mechanism, kinetics, role of microorganisms and effect of different leaching parameters are still under debate. A short review of the concepts that are central to the discussions presented in this investigation is given in the following sections. 1.1. Galvanic dissolution of chalcopyrite The effect of impurities on the dissolution of chalcopyrite is a relatively old observation (Dutrizac and MacDonald, 1973). For sulphide minerals that are in contact with each other, the mineral with higher rest potential acts as a cathode and is preserved, while the mineral with lower rest potential acts as an anode and is oxidised (Mehta and Murr, 1983). For chalcopyrite, pyrite gives rise to increased copper release as it acts as the cathode and chalcopyrite serves as the anode and is oxidised. The main reaction for chalcopyrite dissolution in ferric/ferrous sulphate systems is the oxidation of chalcopyrite by ferric ions:

10

M. Khoshkhoo et al. / Minerals Engineering 100 (2017) 9–16

CuFeS2 þ 4Fe3þ ! Cu2þ þ 5Fe2þ þ 2S0

ð1Þ

Being electrochemical in nature, this reaction can be expressed as an anodic half-cell reaction (Eq. (2)) and a cathodic half-cell reaction (Eq. (3)):

CuFeS2 ! Cu2þ þ Fe2þ þ 2S0 þ 4e

ð2Þ

4Fe3þ þ 4e ! 4Fe2þ

ð3Þ

The limiting step is reported to be the reduction of ferric iron which is slow on the surface of chalcopyrite due to presence of hindering layers (Dixon et al., 2008). When pyrite and chalcopyrite are in contact, the iron reduction takes place on the pyrite surface, which is much faster. As a result, the total dissolution rate is increased.

h  i 4þ 3nCuFeS2ðsolidÞ þ ð12n  12ÞFe3þ  6 S2 n ðaqÞ ! 4ðCuFeÞ

1.2. Reductive leaching mechanism Reductive leaching of chalcopyrite generally refers to dissolution of chalcopyrite at low redox potential. The redox potential of a system containing ferric and ferrous ions is mainly determined by the ratio of free ferric to ferrous ions, according to the Nernst equation (Eq. (4)):

Eh ¼ E þ

dissolution rate is optimal. This range is reported to be around 600–680 mV vs SHE (Koleini et al., 2011) which is considered as a low redox potential level in chalcopyrite leaching systems. The reductive leaching mechanism has also been described by another model based on evolution of surface layers measured by X-ray photoelectron spectroscopy (XPS), time of flight secondary ion mass spectrometry (ToF-SIMS) and scanning electron microscopy (SEM) (Harmer et al., 2006). In this model, it is suggested that chalcopyrite is partially oxidised by ferric ions in the first step, resulting in a polysulphide phase and release of cupric and ferrous ions (Eq. (8)). The second step consists of reduction of the polysulphide intermediate by ferrous ions to sulphide (Eq. (9)) and finally, the sulphide is oxidised with ferric ions forming elemental sulphur accompanied by additional release of cupric and ferrous ions into the solution (Eq. (10)). This model also requires a relatively low redox potential to satisfy the conditions for oxidation and reduction reactions.

R  T fFe3þ g ln n  F fFe2þ g

ð4Þ

where E° is the standard electrode potential, R is the gas constant, n is the charge number, F is Faraday’s constant, and {Fe3+} and {Fe2+} are activities of ferric and ferrous ions, respectively. From chalcopyrite oxidation by ferric ions (Eq. (1)), it seems that higher redox potentials, i.e. higher ferric to ferrous ions concentrations, should favour its dissolution. This is not what is usually observed and instead, dissolution increases with increasing redox potential and then decreases after a critical potential is reached (Kametani and Aoki, 1985). A two-step model based on thermodynamic calculations was proposed to support this observation (Hiroyoshi et al., 2000). The model suggests that in the first step chalcopyrite is reduced by ferrous ions in the presence of cupric ions to form chalcocite (Eq. (5)). In the second step, chalcocite which is more amenable to leaching than chalcopyrite is oxidised by ferric ions or oxygen according to Eqs. (6) and (7).

CuFeS2 þ 3Cu2þ þ 3Fe2þ ! 2Cu2 S þ 4Fe3þ

ð5Þ

2Cu2 S þ 8Fe3þ ! 4Cu2þ þ 2S0 þ 8Fe2þ

ð6Þ

2Cu2 S þ 8Hþ þ 2O2 ! 4Cu2þ þ 2S0 þ 4H2 O

ð7Þ

This model suggests that the redox potential has to be low enough for reduction of chalcopyrite in the first step and high enough for oxidation of chalcocite in the second. Consequently, there is a window of oxidation potential in which the chalcopyrite

2þ þ ð3n  4ÞCu2þ ðaqÞ þ ð15n  16ÞFeðaqÞ

ðsolidÞ

where n P 2

ð8Þ

h  i þ þ ð12n  12ÞFe2þ 4ðCuFeÞ4þ  6 S2 ðaqÞ þ ð12n  12ÞHðaqÞ n ðsolidÞ h i ! 4ðCuFeÞ4þ  ð12n  12ÞHþ  6nðS2 Þ þ ð12n  12ÞFe3þ ðaqÞ ðsolidÞ

where n P 2

ð9Þ

0 ½4ðCuFeÞ4þ  ð12n  12ÞHþ  6nðS2 ÞðsolidÞ þ ð12nÞFe3þ ðaqÞ ! 6nSðsolidÞ 2þ þ þ 4Cu2þ ðaqÞ þ ð12n þ 4ÞFeðaqÞ þ ð12n  12ÞHðaqÞ

where n P 2 ð10Þ

In the present investigation, results from leaching of a pure concentrate as well as a pyritic concentrate performed under strictly controlled redox potential conditions are presented and the interaction of galvanic dissolution with reductive leaching mechanism is discussed. 2. Materials and methods 2.1. Aitik chalcopyrite concentrate (high purity concentrate) A high purity chalcopyrite concentrate from the Boliden Mineral AB owned Aitik mine in Sweden was used. The concentrate was washed with ethanol and water, dried by acetone rinsing at room temperature, divided into 15 g lots, sealed in plastic containers flushed with nitrogen and kept at 4 °C. The concentrate was ground in a ring mill for 15 s immediately before addition to the leaching reactors. This procedure was carried out to assure presence of fresh chalcopyrite surfaces and comparatively similar initial surface characteristics in all the experiments. The reproducibility of the grinding procedure was tested in triplicates giving a d50 of 18 ± 1 lm, a d80 of 42 ± 3 lm and a specific surface area of

Table 1 Chemical and mineralogical composition of the concentrate used in leaching experiments. Chemical composition (%)

Aitik (ground) Aitik (38–53 lm) Kristineberg (ground) Kristineberg (38–53 lm) Aitik Pyrite Conc. a

Calculated mineralogy (%)

Cu

Fe

S

Zn

Pb

SiO2

CuFeS2

FeS2

ZnS

PbS

32.4 31.8 23.6 21.6 0.04

30.9 30.2 34.7 33.8 36.4

34.1 35.2 37.5 39.9 39.6

0.11 0.05 2.10 1.7 <0.01

0.05 0.01 0.70 0.58 <0.01

0.53 0.45 1.10 0.68 23.0a

94.0 92.0 68.0 62.4 0.12

4.00 6.00 26 32.4 76.0

0.16 0.07 3.20 2.53 0.01

0.06 0.01 0.80 0.67 <0.01

Total gangue including aluminosilicate minerals (mostly kyanite) for pyrite concentrate.

M. Khoshkhoo et al. / Minerals Engineering 100 (2017) 9–16

11

mineral in the concentrate (Table 1 and Figs. 1 and 2a). Part of the Aitik concentrate that had been stored for two years in the original packing at room temperature was used to investigate the effect of ageing. The aged concentrate was washed with ethanol and water, wet sieved, and the particle size fraction between 38 and 53 lm was dried with acetone rinsing and aliquots sealed in containers flushed with nitrogen. The sealed containers were kept in a refrigerator (4 °C) until used. The fraction’s chemical analysis and calculated mineralogy are given in Table 1 and its surface area was 0.38 ± 0.01 m2/g. 2.2. Kristineberg chalcopyrite concentrate (pyritic concentrate) The regular chalcopyrite concentrate from Kristineberg mine, Boliden Mineral AB contained higher amount of pyrite compared to the Aitik concentrate (Table 1 and Fig. 1). SEM/EDS analyses showed that in some grains pyrite and chalcopyrite co-existed (Fig. 2b). Leaching experiments were performed on the ground concentrate as well as the aged 38–53 lm fractions (approximately two years old). The preparation of the concentrate was similar to the Aitik concentrate. For producing fresh surfaces, the concentrate was ground for 10 s in the ring mill before each experiment giving a d50 of 18 ± 1 lm, a d80 of 45 ± 1 lm and a specific surface area of 0.75 ± 0.02 m2/g. The specific surface area of the 38–53 lm fractions were 0.35 ± 0.01 m2/g. Fig. 1. XRD patterns of Aitik and Kristineberg concentrate. Abbreviations: C: chalcopyrite, P: pyrite, G: ganuge.

2.3. Aitik pyrite concentrate The pyrite concentrate originating from the Aitik mine was obtained from tailings of a chalcopyrite flotation at the Boliden plant and acid washed for removal of the acid soluble gangues. The amount of pyrite in the concentrate calculated based on the chemical analyses after the acid washing stage was around 74% (Table 1). The total amount of the gangue minerals, mostly as kyanite (Al2SiO5), was around 23%. 2.4. Leaching experiments under controlled redox potential conditions

Fig. 2. Backscattered electron image of (a) Aitik concentrate and (b) Kristineberg concentrate showing chalcopyrite (C), pyrite (P) and gangue minerals (G).

0.90 ± 0.03 m2/g. X-ray diffraction (XRD) and scanning electron microscopy coupled with energy-dispersive X-ray spectroscopy (SEM/EDS) analyses showed that chalcopyrite was the only copper

Batch leaching experiments were performed in 2 l baffled reactors with a working volume of 1.25 l. Heating was provided by an electric plate controlled at 80 °C (unless stated otherwise) by an Ebro GFX 460 Fuzzy Controller. The pH of the solution was regularly monitored and controlled at 1.5 by manual addition of 5 M sulphuric acid during the experiments. The initial iron concentration (1 g/l (0.02 M) and in some experiments 5 g/l (0.09 M)) was supplied by addition of ferrous sulphate. When applicable, the initial concentration of copper was adjusted by addition of copper sulphate. Stirring was provided by an overhead paddle stirrer at 400 rpm in all experiments. The reactors were equipped with condensers to minimise evaporation. After the temperature and redox potential were set at the desired level, 12.5 g of concentrate (1% (wt/vol) solid content) was added to the reactor. The redox potential of the solution was continuously measured using a platinum electrode with a Ag/AgCl reference electrode (Metrohm) containing 3 M KCl solution as the electrolyte. Unless otherwise specified, all reported redox potentials are versus the Ag/AgCl/3 M KCl reference electrode. The electrodes were checked with a reference solution at frequent intervals and were replaced with a fresh and clean electrode at least once a day. The redox potential was adjusted to the set value by automatic addition of a dilute hydrogen peroxide solution (1.5% v/v). The program for setting the redox potential level and controlling the addition of the oxidising agent was written in LabView software (National Instruments). In experiments controlled at a low redox potential level (420–450 mV), nitrogen was sparged into the reactors, minimising the oxidation of ferrous to ferric ions by air, and thus avoiding a loss of redox

12

M. Khoshkhoo et al. / Minerals Engineering 100 (2017) 9–16

potential control. The nitrogen was of chemical quality provided by AGA with 65 ppm of oxygen and moisture. The described setup successfully controlled the redox potential ±5 mV of the set values during the relatively long leaching periods in all the experiments. One experiment was carried out at a lower redox potential level (350 mV) where the redox was controlled by automatic addition of a dilute sodium sulphite solution. The reactors were regularly sampled and the pulp volume was monitored by weighing before sampling and adjusted if necessary. The samples were filtered and the remaining solids were returned to the reactors. The amount of dissolved copper removed during sampling was accounted for in the recovery calculations. The pulp was filtered at the end of the experiments, washed, dried at room temperature and saved for further analysis. Four selected experiments were performed in duplicates and the maximum standard deviation in copper recovery was found to be 1.7. 2.5. Calculation of activation energies Initial activation energies at low (420 mV) and high (620 mV) redox potentials were calculated by plotting the Arrhenius equation (Eq. (11)) where k is the leaching rate (mg/h), A is the pre-exponential factor, Ea is the activation energy, R is the universal gas constant and T is temperature (K).

ln k ¼ ln A 

Ea RT

ð11Þ

Rates from the initial stage of leaching experiments (recoveries below 10%) at different temperatures were extracted from the leaching data by fitting a straight line to at least four data points, resulting in R2 of 0.99 or better for all the fits. The natural logarithms of rates (in mg/h) were plotted against the reciprocal of the temperature (K). The resulting Arrhenius plots were fitted by a linear equation and the slopes of the line (Ea/R) were used for activation energy calculations. 2.6. Analyses Dissolved iron and copper in the solution samples were analysed by atomic absorption spectroscopy (AAS; Perkin Elmer, Analyst 100). A CILAS 1064 Liquid apparatus and a Micromeritics Flowsorb II 2300 were employed for particle size distributions and surface area measurements, respectively. The phases and the minerals present in the concentrate and residues were examined by a PANalytical Empyrean X-ray powder diffractometer. Patterns were collected between 2h angles of 5–90° at Cu Ka radiation of 45 kV, 40 mA. The software HighScore Plus supplemented with COD 2013 database was utilised to fit and analyse the diffractograms. A backscattered electron detector on a Ziess (MERLIN HR) scanning electron microscopy (SEM) was used to produce SEM photographs and energy dispersive X-ray spectroscopy (EDS) analyses. The acceleration voltage was set to 20 keV and the emission current was 1.0 nA.

Fig. 3. Effect of redox potential in leaching of freshly ground Kristineberg concentrate. Leaching conditions: T = 80 °C, pH = 1.4–1.5, initial iron: 1 g/l, initial copper: 0 g/l. Solid content (S.C.) in the longer experiment at 620 mV was 0.3% (w/v).

with a lower solid content (4 g instead of 12.5 g) so the redox potential could be accurately controlled over a longer leaching period. In leaching of the relatively pure Aitik concentrate, in absence of pyrite, the effect of reductive leaching was not apparent and the recoveries became more or less similar at low and high redox potentials (Fig. 4). When pyrite was added to the freshly ground Aitik concentrate and the experiment were repeated at 420 mV, the leaching was improved from 50% to around 65% after 24 h of leaching (Fig. 5; results from leaching of the Kristineberg concentrate at 420 mV is also given for comparison). The amount of pyrite and chalcopyrite concentrates in this experiment reproduced the pyrite:chalcopyrite weight ratio in the Kristineberg concentrate (0.38) while keeping the total solid content at around 1% (9.04 g Aitik concentrate mixed with 3.71 g pyrite concentrate). The difference between copper recoveries in Kristineberg leaching experiment and pyrite added Aitik experiment could be due to the differences in the pyrite source, the presence of pyrite in the same grains with chalcopyrite in the Kristineberg concentrate (Fig. 2b) as well as a different particle size distribution of chalcopyrite in the ground Kristineberg compared to that in the Aitik concentrate. The particle size distribution of the freshly ground concentrates were almost similar, but the Kristineberg concentrate contained pyrite as well as chalcopyrite and it is possible that their

3. Results and discussion 3.1. Effect of pyrite on reductive leaching In agreement with the common understanding of the chalcopyrite reductive leaching mechanism (Hiroyoshi et al., 2000; Kametani and Aoki, 1985; Sandström et al., 2005) copper recovery in leaching of the fresh Kristineberg concentrate performed at low redox potential (420 mV) was significantly higher than that obtained at high redox potential (620 mV) after 24 h of leaching (Fig. 3). Due to high reactivity of the pyrite content in the Kristineberg concentrate at 620 mV, the experiment was repeated

Fig. 4. Effect of redox potential in leaching of freshly ground Aitik concentrate. Leaching conditions: T = 80 °C, pH = 1.4–1.5, initial iron: 1 g/l, initial copper: 0 g/l.

M. Khoshkhoo et al. / Minerals Engineering 100 (2017) 9–16

Fig. 5. Effect of pyrite addition in leaching of freshly ground Aitik concentrate. Leaching conditions: redox potential = 420 mV, T = 80 °C, pH = 1.4–1.5, Initial iron = 1 g/l, initial copper = 0 g/l. Results from leaching of freshly ground Kristineberg concentrate are also given for comparison.

respective particle size distribution was different from the total particle size distribution. These results generally indicated that in absence of pyrite the reductive leaching mechanism at low redox potentials would not improve the leaching efficiencies, compared to high redox potential experiments. A review of the most cited research articles suggesting higher chalcopyrite dissolution rates within a low redox potential window with long leaching experiments performed in ferric/ferrous sulphate media revealed that most studies actually used a pyritic concentrate (sometimes by addition of extra pyrite) with a pyrite:chalcopyrite weight ratio of P0.15 (Ahmadi et al., 2011; Córdoba et al., 2008; Gericke et al., 2010; Kametani and Aoki, 1985; Koleini et al., 2011; Petersen and Dixon, 2006; Sandström et al., 2005). It is difficult to explain the effect of pyrite in accelerating the reductive leaching of chalcopyrite using the model proposed by Hiroyoshi et al. (2000) as the role of galvanic interaction is to provide a surface on which the limiting step of ferric to ferrous reduction occurs faster compared to on the chalcopyrite surface (Eq. (3)). In the Hiroyoshi et al. model, ferric to ferrous reduction occurs during chalcocite oxidation (Eq. (6)) which regardless of pyrite presence or absence is faster than chalcopyrite oxidation. Consequently, the dissolution rate of chalcopyrite in the absence of pyrite at low redox potentials should still be higher compared to the experiments at high redox potentials. However, this was not observed in the long leaching experiments under controlled redox potential conditions in this investigation (Fig. 4). Instead, in the absence of pyrite the dissolution rates were comparable at low and high redox potentials after 72 h of leaching. It is easier to explain these results by a model similar to the one proposed by Harmer et al. (2006) whereby the first step is the partial oxidation of chalcopyrite by ferric ions (Eq. (8)). This equation, on one hand, describes that in the absence of pyrite at low and high redox potentials similar copper dissolution is obtained. On the other hand, it supports the significant improvement of copper dissolution at low redox potential conditions in the presence of pyrite by providing a surface on which reduction of ferric to ferrous takes place faster.

13

generally rose with an increase in the redox potential (Fig. 6; experiments with 1.5 g/l initial copper). A similar trend was observed in experiments without addition of initial copper that included one experiment at 350 mV (Fig. 7; see Section 3.3 for discussion of the effect of the initial copper concentration). Combining the results in Figs. 6 and 7 suggested that at a redox potential between 350 and 450 mV, the copper recoveries were nearly insensitive to redox potential variation but increased when the redox potential was >500 mV. Copper dissolution from the aged pyritic Kristineberg concentrate (38–53 lm fractions) was also greater at higher redox potential (Fig. 8). The leaching results from fresh and aged concentrates in this investigation cannot be compared with each other directly, due to their different particle size distribution. However, it has been previously shown that when an aged concentrate is leached, the copper dissolution efficiencies are lower compared to leaching of a concentrate with fresh surfaces under similar leaching conditions (Khoshkhoo et al., 2014b). Atmospheric oxidation of chalcopyrite (ageing) produces compounds such as iron–oxyhydroxides and basic iron–sulphates on the chalcopyrite surface (Brion, 1980; Buckley and Woods, 1984; Hackl et al., 1995; Harmer et al., 2006; Klauber et al., 2001; Mikhlin et al., 2004). Our previous XPS studies also showed the presence of such compounds on the surfaces of the same aged chalcopyrite samples from Aitik (Khoshkhoo et al., 2014a) and

Fig. 6. Effect of redox potential in aged Aitik concentrate (38–53 lm) leaching. Leaching conditions: T = 80 °C, pH = 1.4–1.5, initial iron: 1 g/l, initial copper: 1.5 g/l.

3.2. Effect of concentrate ageing in leaching behaviour The effect of redox potential on leaching behaviour completely changed when the aged concentrates were used. The copper recoveries from the aged Aitik concentrate (38–53 lm fractions)

Fig. 7. Effect of redox potential in aged Aitik concentrate (38–53 lm) leaching. Leaching conditions: T = 80 °C, pH = 1.4–1.5, initial iron: 1 g/l, initial copper: 0 g/l.

14

M. Khoshkhoo et al. / Minerals Engineering 100 (2017) 9–16

Fig. 8. Effect of redox potential in aged Kristineberg concentrate (38–53 lm) leaching. Leaching conditions: T = 80 °C, pH = 1.4–1.5, initial iron: 1 g/l, initial copper: 0 g/l.

Kristineberg (Khoshkhoo et al., 2014b). It is interesting to observe how the presence of these compounds on the surface, although in minute amounts, can change the leaching response in regards to the redox potential conditions. This also implies that the mechanism of leaching is probably different in presence of ageing products. The activation energies of both aged (38–53 lm fractions) and freshly ground Aitik concentrate during the initial leaching periods (recoveries below 10%) were >40 kJ/mol and indicated a chemically controlled reaction (Habashi, 1999) in agreement with many other reports in the literature (Li et al., 2013). However, the activation energy of the aged concentrate leached at 420 mV was much less than that for the fresh concentrate at 420 mV (76.5 vs. 101.9 kJ/mol) as well as less than that for the aged concentrate leached at 620 mV (76.5 vs. 90.9 kJ/mol; Fig. 9, Table 2). This indicates that the leaching of the aged concentrate at low redox potential became less sensitive to the temperature variation, due to the presence of atmospheric oxidation products on the surface of the mineral. At high redox potential, leaching of the aged concentrate, although considerably slower, was equally temperature sensitive as the fresh concentrate (90.9 vs. 91.8 kJ/mol). It is not completely clear why the ageing products only decreased the activation energy at low redox potential, but it can be speculated that the higher activity of ferric ions (main oxidant) at high redox potential made the reaction equally temperature sensitive in spite of the aged surfaces. Unfortunately, calculations of activation energies at later stages of leaching proved to be complicated. The main problem was that in experiments at low temperature and redox potentials where the rate of reactions were too slow (and decreasing as the leaching progressed) the control of the redox potential was lost before significant levels of recoveries were reached. Ageing would not be a concern in industrial scale tank leaching/ bioleaching as the ore is treated in successive stages of grinding, flotation and leaching (Khoshkhoo et al., 2014b). However, heap leaching might take months to years and as a result, atmospheric oxidation can be a factor in leaching efficiency. It is also important that laboratory investigations take into account the use of aged chalcopyrite concentrates. It is normally impossible to identify ageing by conventional XRD and SEM measurements while the availability of surface analytical methods such as XPS and auger electron spectroscopy (AES) is also limited. In addition, atmospheric oxidation is a slow but progressive process that can occur in the presence of minute amounts of air and moisture and its extent can be different depending on the age of samples and humidity. It has been shown that after three months of ageing the concentrate behaved differently under similar leaching conditions (Khoshkhoo et al., 2014b). Thus, research laboratories should

Fig. 9. Arrhenius plots for initial dissolution of (a) aged Aitik concentrate (38–53 lm) and (b) freshly ground Aitik concentrate leached at 420 mV and 620 mV.

Table 2 List of experiments performed for activation energy calculations. Experiment

Concentrate

Redox potential (mV)

Temp. (°C)

Ea (kJ/mol)

A-420-50 A-420-70 A-420-90 A-620-30 A-620-60 A-620-80

Aged Aitik 38–53 lm

420 420 420 620 620 620

50 70 90 30 60 80

76.5

F-420-40 F-420-60 F-420-80 F-620-40 F-620-60 F-620-80

Freshly ground Aitik

420 420 420 620 620 620

40 60 80 40 60 80

101.9

90.9

91.8

take care to ensure that the concentrate surface characteristics are similar and of sufficient quality in all experiments. Unfortunately, the effect of ageing is either underestimated or completely ignored in much of the published literature. This might be an important factor causing the wide and sometimes contradictory results regarding different aspects of chalcopyrite leaching.

M. Khoshkhoo et al. / Minerals Engineering 100 (2017) 9–16

3.3. Effect of initial copper concentration Two experiments were performed in the absence and presence of 1.5 g/l of initial copper (equivalent to 65% recovery) to examine if initial cupric ions could possibly have a positive effect on improving the dissolution rates. The experiments were performed on freshly ground Kristineberg concentrate due to its higher reactivity, the redox potential was controlled at 420 mV and the temperature was 80 °C. The results were similar in both experiments indicating that the presence of cupric ions did not enhance the dissolution rates (Fig. 10). In another series of experiments with 38–53 lm fraction of the aged Aitik concentrate the effect of initial copper at both low (420 mV) and high (620 mV) redox potentials was investigated. It was again observed that the leaching rates were higher at high redox potential experiments (see Section 3.2). At low redox potential a consistent trend versus variation of copper concentrations was not observed (Fig. 11). At high redox, it was still difficult to find a relation, but generally it seems that the recoveries were slightly lower in presence of copper (27% vs. 35% after 72 h). The reason for this behaviour could have been due to the aged surfaces of the concentrate which might have affected the consistency of

Fig. 10. Effect of initial copper concentration in leaching of freshly ground Kristineberg concentrate. Leaching conditions: redox potential = 420 mV, T = 80 °C, pH = 1.4–1.5, initial iron = 1 g/l.

Fig. 11. Effect of initial copper concentration in leaching of aged Aitik concentrate (38–53 lm). Legend: First number shows the redox potential (mV) and the second number is the initial copper concentration (g/l). Other leaching conditions: T = 80 °C, pH = 1.4–1.5, initial iron = 1 g/l.

15

the results. Nevertheless, it seems that the effect of cupric ion concentration in leaching efficiencies is not that profound to be of concern or interest in any possible industrial application. 3.4. Hindered dissolution: a few remarks The formation of iron–oxyhydroxides on chalcopyrite surfaces as a result of atmospheric oxidation is a well–known phenomenon (Brion, 1980; Buckley and Woods, 1984; Hackl et al., 1995; Harmer et al., 2006; Klauber et al., 2001; Mikhlin et al., 2004). Upon longer exposure and in the presence of moisture, partial or total hydrolysis of iron–oxyhydroxides results in formation of basic ferric–sulphate (Brion, 1980; Parker, 2008). It is reasonable to assume that this process would occur in sulphate leaching systems at an accelerated rate. The presence of iron–oxyhydroxides and basic iron–sulphates has been reported in leaching investigations in sulphate systems and atmospheric pressure (Farquhar et al., 2003; Khoshkhoo et al., 2014a; Parker, 2008). They are also observed on surface of samples leached in more acidic conditions (Hackl et al., 1995) and in chloride media (Acres et al., 2010; Li et al., 2010). Formation of iron-oxyhydroxides on the surface leads to formation of a phase rich in sulphur and copper below the oxidised surface, typically regarded as polysulphide or metaldeficient sulphide. Some investigations considered this phase as the cause for hindered dissolution (Hackl et al., 1995; Harmer et al., 2006) while others suggested that this phase can only be a transient intermediate and has no role in hindering the dissolution (Holmes and Crundwell, 2013; Klauber, 2008; Parker et al., 2003). Results from the current investigation are inconclusive for proposing a definitive cause for hindered dissolution. However, the suggestion that iron–oxyhydroxides could be the precursor for passivation or hindered dissolution is supported by experiments showing lower leaching rates in aged concentrates from the outset even when the conditions that fresh chalcopyrite would react rapidly were used. The hypothesis that iron-oxyhydroxides trigger hindered dissolution is also supported by XPS studies in our previous investigation where the presence of iron-oxyhydroxides (mixed with elemental sulphur) was related to the slow dissolution (Khoshkhoo et al., 2014a). It was found that elemental sulphur was rigidly bound to the surface such that the XPS measurements at room temperature, where sublimation of sulphur must occur under ultra-high vacuum in the XPS analysing chamber, did not result in removal of sulphur from the surface. It can be proposed that as the dissolution of chalcopyrite in sulphate media is initiated, elemental sulphur is gradually trapped in already and newly formed iron–oxyhydroxides (and subsequently basic ironsulphates) resulting in a multicomponent layer which hinders the dissolution. If this is the case, then it seems that hindered dissolution cannot be avoided in conventional sulphate leaching systems, especially in bioleaching systems at high redox potentials and extended leaching durations where iron–oxyhydroxides would have enough time to form and become consolidated on the surface. One possible solution would be to leach under suitable conditions such that chalcopyrite dissolution reaches an acceptable level before its surface is extensively covered. Suitable conditions are fresh surfaces, temperature >80 °C, low redox potential and presence of pyrite. Such an experiment is shown in Fig. 3 (freshly ground Kristineberg leached at 420 mV) where the recovery reached above 70% in eight hours, after which it started to be hindered. Dixon et al. (2008) and others have shown this can be further improved with higher amounts of pyrite and probably by higher solid contents to increase the collisions between pyrite and chalcopyrite particles.

16

M. Khoshkhoo et al. / Minerals Engineering 100 (2017) 9–16

4. Conclusions Controlled redox potential leaching experiments showed that the reductive leaching mechanism would be beneficial in the presence of an active galvanic interaction. Experiments using a pyritic concentrate resulted in higher recoveries at low redox potential while the recoveries were similar at low and high redox potentials using a relatively pure concentrate in absence of pyrite. This behaviour was shown to be in better agreement with the alternative reductive leaching model that suggests chalcopyrite surfaces undergo a first step of partial oxidation before the reduction reactions occur. This model accounts for the effect of pyrite in the reductive leaching mechanism. It was also shown that the reductive leaching was not valid at all for aged concentrates. Aged concentrates (pyritic and pure) gave steadily increasing recoveries with increased redox potential. The air oxidation products, mainly iron-oxyhydroxide, on the surfaces of the minerals were found to cause such behaviour. Chalcopyrite ageing can have drastic effects in heap leaching processes where the ore is exposed for extended periods of time. The wide and inconsistent chalcopyrite leaching results reported in the literature could also be partly due to using of aged samples. The effect of initial copper concentration had no influential effect in the leaching rates for possible industrial processes. It was also proposed that iron–oxyhydroxides are the precursor of hindered dissolution in sulphate media where their formation entraps surface elemental sulphur and other surface species resulting to an inevitable formation of a consolidated phase on the surface. It was shown that when suitable conditions are met, high copper recoveries can be obtained before the surface is extensively hindered. Acknowledgements The authors wish to express their gratitude for financial support from the Swedish Research Council Formas (contract number 2292009-657). Boliden Mineral AB is highly appreciated for providing financial support and mineral samples. Dr. Roland Hostettler is also gratefully acknowledged for his invaluable assistance in improving the redox potential control program. References Acres, R., Harmer, S., Beattie, D., 2010. Synchrotron XPS studies of solution exposed chalcopyrite, bornite, and heterogeneous chalcopyrite with bornite. Int. J. Miner. Process. 94, 43–51. http://dx.doi.org/10.1016/j.minpro.2009.11.006. Ahmadi, A., Schaffie, M., Petersen, J., Schippers, A., Ranjbar, M., 2011. Conventional and electrochemical bioleaching of chalcopyrite concentrates by moderately thermophilic bacteria at high pulp density. Hydrometallurgy 106, 84–92. http:// dx.doi.org/10.1016/j.hydromet.2010.12.007. Brion, D., 1980. Etude par spectroscopie de photoelectrons de la degradation superficielle de FeS2, CuFeS2, ZnS et PbS a l’air et dans l’eau. Appl. Surf. Sci. 5, 133–152. http://dx.doi.org/10.1016/0378-5963(80)90148-8. Buckley, A., Woods, R., 1984. An X-ray photoelectron spectroscopic study of the oxidation of chalcopyrite. Aust. J. Chem. 37, 2403. Córdoba, E.M., Muñoz, J.A., Blázquez, M.L., González, F., Ballester, A., 2008. Leaching of chalcopyrite with ferric ion. Part II: effect of redox potential. Hydrometallurgy 93, 88–96. http://dx.doi.org/10.1016/j.hydromet.2008.04.016. Debernardi, G., Carlesi, C., 2013. Chemical-electrochemical approaches to the study passivation of chalcopyrite. Miner. Process. Extr. Metall. Rev. 34, 10–41. http:// dx.doi.org/10.1080/08827508.2011.623745. Dixon, D.G., Mayne, D.D., Baxter, K.G., 2008. Galvanox – a novel galvanicallyassisted atmospheric leaching technology for copper concentrate. Can. Metall. Quart. 47, 327–336. http://dx.doi.org/10.1179/000844308794408317.

Dreisinger, D., 2006. Copper leaching from primary sulfides: options for biological and chemical extraction of copper. Hydrometallurgy 83, 10–20. http://dx.doi. org/10.1016/j.hydromet.2006.03.032. Dutrizac, J.E., MacDonald, R.J.C., 1973. The effect of some impurities on the rate of chalcopyrite dissolution. Can. Metall. Quart. 12, 409–420. Farquhar, M.L., Wincott, P.L., Wogelius, R.A., Vaughan, D.J., 2003. Electrochemical oxidation of the chalcopyrite surface: an XPS and AFM study in solution at pH 4. Appl. Surf. Sci. 218, 34–43. Gericke, M., Govender, Y., Pinches, A., 2010. Tank bioleaching of low-grade chalcopyrite concentrates using redox control. Hydrometallurgy 104, 414– 419. http://dx.doi.org/10.1016/j.hydromet.2010.02.024. Habashi, F., 1999. Kinetics of Metallurgical Processes. Metallurgie Extractive Quebec, Quebec City. Hackl, R.P., Dreisinger, D.B., Peters, E., King, J.A., 1995. Passivation of chalcopyrite during oxidative leaching in sulfate media. Hydrometallurgy 39 (25–48), 25. http://dx.doi.org/10.1016/0304-386X(95)00023-A. Harmer, S.L., Thomas, J.E., Fornasiero, D., Gerson, A.R., 2006. The evolution of surface layers formed during chalcopyrite leaching. Geochim. Cosmochim. Acta 70, 4392–4402. http://dx.doi.org/10.1016/j.gca.2006.06.1555. Hiroyoshi, N., Miki, H., Hirajima, T., Tsunekawa, M., 2000. A model for ferrouspromoted chalcopyrite leaching. Hydrometallurgy 57, 31–38. http://dx.doi.org/ 10.1016/S0304-386X(00)00089-X. Holmes, P.R., Crundwell, F.K., 2013. Polysulfides do not cause passivation: results from the dissolution of pyrite and implications for other sulfide minerals. Hydrometallurgy 139, 101–110. http://dx.doi.org/10.1016/j.hydromet.2013. 07.006. Kametani, H., Aoki, A., 1985. Effect of suspension potential on the oxidation rate of copper concentrate in a sulfuric acid solution. Metall. Mater. Trans. B 16, 695– 705. http://dx.doi.org/10.1007/BF02667506. Khoshkhoo, M., Dopson, M., Shchukarev, A., Sandström, Å., 2014a. Chalcopyrite leaching and bioleaching: an X-ray photoelectron spectroscopic (XPS) investigation on the nature of hindered dissolution. Hydrometallurgy 149, 220–227. http://dx.doi.org/10.1016/j.hydromet.2014.08.012. Khoshkhoo, M., Dopson, M., Shchukarev, A., Sandström, Å., 2014b. Electrochemical simulation of redox potential development in bioleaching of a pyritic chalcopyrite concentrate. Hydrometallurgy 144–145, 7–14. http://dx.doi.org/ 10.1016/j.hydromet.2013.12.003. Klauber, C., 2008. A critical review of the surface chemistry of acidic ferric sulphate dissolution of chalcopyrite with regards to hindered dissolution. Int. J. Miner. Process. 86, 1–17. http://dx.doi.org/10.106/j.minpro.2007.09.003. Klauber, C., Parker, A., Bronswijk, W.v., Watling, H., 2001. Sulphur speciation of leached chalcopyrite surfaces as determined by X-ray photoelectron spectroscopy. Int. J. Miner. Process. 62 (65–94), 65. http://dx.doi.org/10.1016/ S0301-7516(00)00045-4. Koleini, S.M.J., Aghazadeh, V., Sandström, A., 2011. Acidic sulphate leaching of chalcopyrite concentrates in presence of pyrite. Miner. Eng. 24, 381–386. http:// dx.doi.org/10.1016/j.mineng.2010.11.008. Li, Y., Kawashima, N., Li, J., Chandra, A.P., Gerson, A.R., 2013. A review of the structure, and fundamental mechanisms and kinetics of the leaching of chalcopyrite. Adv. Colloid Interface Sci. 197–198, 1–32. http://dx.doi.org/ 10.1016/j.cis.2013.03.004. Li, J., Kawashima, N., Kaplun, K., Absolon, V.J., Gerson, A.R., 2010. Chalcopyrite leaching: the rate controlling factors. Geochim. Cosmochim. Acta 74, 2881– 2893. http://dx.doi.org/10.1016/j.gca.2010.02.029. Marsden, J., 2007. Medium-temperature pressure leaching of copper concentrates Part 1: chemistry and initial process development. Miner. Metall. Process. 24, 193–204. Mehta, A.P., Murr, L.E., 1983. Fundamental studies of the contribution of galvanic interaction to acid-bacterial leaching of mixed metal sulfides. Hydrometallurgy 9, 235–256. http://dx.doi.org/10.1016/0304-386X(83)90025-7. Mikhlin, Y.L., Tomashevich, Y.V., Asanov, I.P., Okotrub, A.V., Varnek, V.A., Vyalikh, D. V., 2004. Spectroscopic and electrochemical characterization of the surface layers of chalcopyrite (CuFeS2) reacted in acidic solutions. Appl. Surf. Sci. 225, 395–409. http://dx.doi.org/10.1016/j.apsusc.2003.10.030. Parker, A., Klauber, C., Kougianous, A., Watling, H.R., van Bronswijk, W., 2003. An Xray photoelectron spectroscopy study of the mechanism of oxidative dissolution of chalcopyrite. Hydrometallurgy 71, 265–276. http://dx.doi.org/ 10.1016/S0304-386X(03)00165-8. Parker, A., 2008. Oxidative Dissolution of Chalcopyrite in Ferric Media: An X-ray Photoelectron Spectroscopy Study PhD dissertation [PhD Thesis]. Curtin University of Technology. Petersen, J., Dixon, D.G., 2006. Competitive bioleaching of pyrite and chalcopyrite. Hydrometallurgy 83, 40–49. http://dx.doi.org/10.1016/j.hydromet.2006.03.036. Sandström, Å., Shchukarev, A., Paul, J., 2005. XPS characterisation of chalcopyrite chemically and bio-leached at high and low redox potential. Miner. Eng. 18, 505–515. http://dx.doi.org/10.1016/j.mineng.2004.08.004.