Lixiviación Calcopirita Con Ozono

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Journal of Industrial and Engineering Chemistry 104 (2021) 333–344

Contents lists available at ScienceDirect

Journal of Industrial and Engineering Chemistry


journal homepage: www.elsevier.com/locate/jiec

Evaluation of ozone as an efficient and sustainable reagent for


chalcopyrite leaching: Process optimization and oxidative mechanism
Jingxiu Wang, Fariborz Faraji ⇑, Ahmad Ghahreman ⇑
Hydrometallurgy and Environmental Laboratory, Robert M. Buchan Department of Mining, Queen’s University, 25 Union Street, Kingston, Ontario K7L 3N6, Canada

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

Article history: Chalcopyrite is the main copper-bearing mineral and is refractory to oxidative leaching. There are many
Received 7 February 2021 investigations to optimize copper extraction from this mineral but most of the processes are energy
Revised 22 August 2021 intensive and environmental threatening. Ozone at low concentration is an efficient reagent for high
Accepted 23 August 2021
extraction of copper from chalcopyrite, which is also ecofriendly. This paper presents a comprehensive
Available online 28 August 2021
investigation on the process optimization for both copper and iron extraction from chalcopyrite using
ozone. The following variables were tested during 48-h leaching periods: temperature, ferric sulfate,
Keywords:
ozone and sulfuric acid concentrations, solid/liquid ratio, and leaching time. Ozone leaching recovered
Ozonation
Sulfidic ores
100% of the copper from chalcopyrite at 25 °C without addition of ferric. The overall reaction kinetics fol-
Copper lowed the shrinking core model by mixed control of diffusion through the product layer and chemical
Leaching mechanism reactions. The proposed leaching mechanism was verified via various characterization techniques.
Ferric sulfate Green chemistry metrics were evaluated, and the optimized process was demonstrated to be environ-
Diffusion inhibition mentally attractive. Ozone leaching appears to have strong potential as a ‘‘green” and technically feasible
method to leach copper from chalcopyrite.
Ó 2021 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights
reserved.

Introduction catalysts) and ultrafine grinding of the feed mineral have been
employed. The most common approach is to leach chalcopyrite
Chalcopyrite (CuFeS2) is among the most abundant at high temperature in dilute sulfuric acid medium with ferric sul-
copper-bearing sulfide minerals, together with chalcocite (Cu2S), fate as the oxidant [10]. The sulfuric acid concentration must be
covellite (CuS), and bornite (Cu5FeS4), and it represents approxi- controlled to prevent iron hydrolysis and precipitation [11].
mately 70% of the Earth’s known copper reserves [1–3]. The industry Although the ferric ion (Fe3+) appears to be responsible for chal-
standard practice of pyrometallurgical smelting can cause environ- copyrite oxidation, it also promotes the formation of a passive
mental problems from sulfur dioxide emissions and is increasingly layer of sulfur on the mineral surface, which reduces leaching
costly in the face of depletion of high-grade ore deposits [4,5]. efficiency. A more efficient and cleaner method for chalcopyrite
Hydrometallurgical processing of copper-bearing minerals has leaching is required.
gained increasing attention because the environmental impacts Common hydrometallurgical leaching oxidants include ferric
are less severe and it is economically feasible for low-grade copper sulfate (Fe2(SO4)3) [10], ferric chloride (FeCl3) [12], hydrogen per-
feeds [6]. oxide (H2O2) [13], ozone (O3) [11], manganese nodules [14], and
Numerous processes have been developed and conditions have various microorganisms [15]. With an oxidation–reduction poten-
been well established for copper dissolution from chalcopyrite ores tial (ORP) of 2.07 V, ozone is a very promising oxidant for removing
and concentrates. However, as the most refractory copper mineral the passive layer during acidic leaching of chalcopyrite [16,17]. It
for hydrometallurgical techniques, chalcopyrite requires highly presents many advantages: since it can be made on-site, no trans-
acidic solutions, elevated pressure and temperature, and thermal portation or storage is required; it is easy to handle; and mainte-
and (or) chemical pretreatments [7,8]. To enhance leaching nance and labor requirements are low [18]. Further, ozone
efficiency, additives (e.g., polar organic reagents [9] and various leaching has been demonstrated to be very effective at room tem-
perature, and the reaction byproduct is O2, thus the impact on the
⇑ Corresponding authors. environment is negligible. Ozone has been used in hydrometallur-
E-mail addresses: f.faraji@queensu.ca (F. Faraji), ahmad.g@queensu.ca gical leaching of zinc from zinc sulfide [19], copper and antimony
(A. Ghahreman). from tetrahedrite [20], stibnite from sulfidic antimony ore [21],

https://doi.org/10.1016/j.jiec.2021.08.036
1226-086X/Ó 2021 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights reserved.
J. Wang, F. Faraji and A. Ghahreman Journal of Industrial and Engineering Chemistry 104 (2021) 333–344

silver from pyrargyrite [22], and gold from concentrates [23,24].


There appears to be a synergistic effect of ferric and ozone in chal-
copyrite leaching [17,25]. Iron dissolution and sulfur oxidation
were 77% and 67%, respectively, under optimum conditions when
ozone was used in combination with Fe3+ in sulfuric acid leaching
of pyrite (FeS2) [23]. Havlik et al. [26] reported up to 80% copper
dissolution from a chalcopyrite concentrate in sulfuric acid solu-
tion with ozone at 22 °C after 40 h.
Depending on the leaching conditions, various reactions might
be involved in copper leaching from chalcopyrite with ozone
[27]. Some oxidize metals whereas others convert sulfide to sulfate
and/or elemental sulfur. Ozone can create favorable conditions to
oxidize chalcopyrite and generate Fe3+ in accordance with reac-
tions (Eq. (1)) and (Eq. (2)) [11].
3CuFeS2 þ 8O3 ! 3CuSO4 þ 3FeSO4 ð1Þ

6FeSO4 þ O3 þ 3H2 SO4 ! 3Fe2 ðSO4 Þ3 þ 3H2 O ð2Þ


Leaching of chalcopyrite by Fe2(SO4)3 and dissolved oxygen Fig. 1. Particle size distribution (PSD) of the chalcopyrite sample.
occurs through reactions (3) and (4), respectively. Regeneration
of Fe3+ obeys reaction (Eq. (5)) [28,29].
prepare leaching solutions, pure (>99% H2SO4) sulfuric acid was
CuFeS2 þ 2Fe2 ðSO4 Þ3 ! CuSO4 þ 5FeSO4 þ 2S0 ð3Þ
diluted to required concentrations in MilliQ deionized water. A sto-
ichiometric amount of iron (III) sulfate hydrate (Fe2(SO4)3) was
CuFeS2 þ 2H2 SO4 þ O2 ! CuSO4 þ FeSO4 þ 2S0 þ 2H2 O ð4Þ added in the solution based on the required concentration.

4FeSO4 þ O2 þ 2H2 SO4 ! 2Fe2 ðSO4 Þ3 þ 2H2 O ð5Þ Experimental design


Thereafter, under a continuous supply of ozone, sulfuric acid
can be formed from elemental sulfur following reaction (Eq. (6)). Experiments were designed to maximize copper extraction
from chalcopyrite, and because the mineral is associated with iron,
S0 þ H2 O þ O3 ! H2 SO4 ð6Þ therefore Fe (ferrous or ferric ions) is also dissolved during this
There is a paucity of information regarding the behavior of process. All experiments were conducted in triplicate and the
ozone in aqueous solutions, and the underlying mechanisms mean ± standard deviation is presented. The effects of six indepen-
responsible for ozone leaching are uncertain and controversial. dent variables on ozone leaching were evaluated:
Therefore, for the first time, this paper presents a detailed opti-
mization of ozone leaching and establishes the ozone leaching 1. Leaching temperature (25, 50, and 85 ℃) at 10.1 ppm ozone in
mechanism. Specific objectives are to: (1) determine the optimal the ozone-oxygen mixture, 10 g/L chalcopyrite, 0.5 M H2SO4,
leaching conditions (temperature, ferric sulfate, ozone, and sulfuric and 24 h leach time.
acid concentrations, solid/liquid (S/L) ratio, and leaching time) in 2. Ferric concentration as ferric sulfate (0, 1, 5, and 10 g/L) at 25 ℃,
ozone-sulfuric acid solution; (2) investigate the effects of Fe3+ 10.1 ppm ozone, 10 g/L chalcopyrite, 0.5 M H2SO4, and 24 h
and (or) oxygen addition on copper, iron, and sulfur recovery; (3) leach time.
study the leaching kinetics of copper in various oxidative leaching 3. Ozone concentration (2.2, 5.3, 10.1, and 19.3 ppm) at 25 ℃, no
systems; (4) evaluate ozone leaching according to green chemistry added ferric, 10 g/L chalcopyrite, 0.5 M H2SO4, and 24 h leach
metrics; and (5) propose a cost-effective, rapid, and efficient time.
approach to improve copper recovery. 4. S/L ratio (3, 5, and 10 g/L chalcopyrite) at 25 ℃, no added ferric,
10.1 ppm ozone, 0.5 M H2SO4, and 24 h leach time.
5. H2SO4 concentration (0.2, 0.5, 1.0, and 2.0 M) at 25 ℃, no added
Experimental
ferric, 10.1 ppm ozone, 5 g/L chalcopyrite, and 24 h leach time.
6. Leach time (1, 3, 5, 8, 24, and 48 h) at 25 ℃, 0 g/L ferric,
Materials
10.1 ppm ozone, 5 g/L chalcopyrite, and 0.5 M H2SO4.

The chalcopyrite sample (Table 1) was crushed, ground, and


To elucidate the ozone leaching mechanism of chalcopyrite dis-
sieved into + 53–75 mm and stored at –25 °C to minimize sulfide
solution and investigate the effect of ferric, two tests were com-
oxidation. Prior to the leaching tests, the sample was dried at
pared at 25, 50, and 85 ℃. Constant conditions were 10 g/L ferric,
40 °C for 24 h. All ore particles were < 75 mm in size (d80 = 60 m
10 g/L chalcopyrite, 0.5 M H2SO4, and 24 h leach time with the fol-
m) with a specific surface area of 653.3 m2/kg (Fig. 1). Quantitative
lowing gas conditions:
X-ray diffractometry (XRD) analysis by HighScore Plus software
showed that the sample comprised of 57.9% CuFeS2, 39.6% FeS2,
1. Ozone-oxygen-ferric test: O2 1 L/min, ozone 10.1 ppm.
and 2.5% quartz (SiO2) [30]. Stoichiometrically, both Fe and S rep-
2. Oxygen-ferric test: O2 1 L/min.
resented 48.8% of the total content, and the remainder was FeS2. To

Table 1
Chemical composition of chalcopyrite sample determined by inductively coupled plasma optical emission spectrometry.

Chalcopyrite elements Cu Fe S Pb Zn Al As Ca K Mg P Na
wt. % 18.2 30 34.3 0.16 0.2 0.5 0.23 1 0.2 0.2 0.2 0.75

334
J. Wang, F. Faraji and A. Ghahreman Journal of Industrial and Engineering Chemistry 104 (2021) 333–344

Experimental apparatus and procedure The particle size of the chalcopyrite powder and leached resi-
dues was determined using a Malvern (United Kingdom) particle
Leaching tests were conducted in a 2-L, five-necked, jacketed size analyzer. The mineralogical composition was determined by
glass reactor with 1-L working volume under mechanical agitation XRD using an X0 Pert Pro Philips powder diffractometer with Cu
of 800 rpm by a glass-bladed overhead stirrer (Fig. 2). Solution Ka radiation (2h range 5–80°, 0.026°/min). Quantitative XRD
temperature was controlled by a jacketed heater regulated with a results were analyzed by using X’Pert HighScore Plus software by
thermocouple connected to an on–off controller. A given amount of the Rietveld method. Morphological characteristics and elemental
chalcopyrite was added to achieve target S/L ratios. Ozone was pro- distributions were investigated by scanning electron microscopy/
duced by an EXT120 ozone generator (Longevity Resources, energy dispersive X-ray spectroscopy (FEI Nova NanoSEM 450,
Canada) using industrial oxygen as the feed gas. The oxygen flow USA). The total sulfur (ST) content was determined by an ELTRA
rate was controlled by a gas flow meter, and the ozone concentra- CS-2000 induction furnace, and elemental sulfur (S0) was mea-
tion was maintained in the ozone-oxygen mixture bubbled into sured with a Netzsch (Germany) STA 449 F3 Jupiter model thermo-
slurry throughout the tests. A reflux condenser prevented evapora- gravimetric analyzer.
tion of the solution, and an ozone destruct unit decomposed excess
ozone to oxygen.
At each sampling time, 10 mL of slurry was collected from the Results and discussion
sampling port (Fig. 2), centrifuged, and filtered through a
0.45 lm syringe membrane filter. The sample volume was replaced Optimization of ozone leaching parameters
with fresh H2SO4 solution. The slurry was vacuum filtered through
a 0.22 lm GF/F glass microfiber filter: the supernatant was diluted Temperature
for measurement of copper and iron concentrations by an MP-AES The highest temperature had the consistently highest copper
4210 microwave plasma-atomic emission spectrometer (Agilent, extraction over a 24-h leach, reaching a maximum of 56.5 ± 1.3%
USA), and the residue was air-dried at room temperature and sent Cu recovery (Fig. 3 (a)). From 8 to 24 h, the extraction rate slowed
for further analysis. The ferrous (Fe2+) concentration was deter- at 50 and 85 °C, but it rapidly increased at 25 °C, such that the
mined by titrating with standardized ceric sulfate to the end point mean Cu recovery after 24 h was higher for 25 °C (45 ± 0.9%) than
using ferroin indicator. Copper and iron concentrations were calcu- 50 °C (36.2 ± 1.1%). A similar pattern was observed for Fe extraction
lated in terms of dissolved metal (%) to compare with residue con- (Fig. 3 (b)), except extraction rates were more rapid for the two
tent for mass balance to ensure accuracy Eq. (7): higher temperatures. The 85 °C treatment yielded a maximum Fe
recovery of 62.7 ± 1.2%. These results differ from a study [11]
CV where chalcopyrite leaching decreased with increasing tempera-
E¼  100 ð7Þ
PM ture, which was ascribed to lower ozone solubility at higher tem-
perature. In the present study, ozone decomposition to hydroxyl
where, E is the Cu/Fe extraction (%), C is the dissolved
radicals may have accelerated at elevated temperature. The stan-
Cu/Fe concentration (g/L), V is the leaching volume (L), P is the
dard hydrogen electrode potential is greater for OH (2.80 V) than
Cu/Fe content in chalcopyrite (%), and M is the mass of added
ozone (2.07 V); therefore, at high temperature, ozone decomposi-
chalcopyrite (g).
tion creates better oxidizing conditions [21], and leaching reac-
tions (3) and (or) (4) are expected to dominate [27]. Indeed, the
highest ORP values were generally seen at 85 °C (Table 2). At
25 °C, the Fe/Cu molar ratio was almost 1 within 24 h, implying
ozone leaching followed reaction (1). Ozone appeared to be more
stable at lower temperatures; therefore, 25 °C was chosen because
the trade-off between energy savings and Cu recovery is consid-
ered satisfactory.

Ferric concentration
Cu recovery was negatively related to ferric concentration: the
highest recovery (45.0 ± 0.9%) was seen in the absence of ferric
and the lowest (18.3 ± 0.8%) was seen at the highest ferric concen-
tration (Fig. 4). At high concentration, ferric stimulates formation
of intermediate products (e.g., SO2–4 ions) that limit leaching. Addi-
tion of 1 g/L ferric may have slightly improved Cu extraction in the
first 3 h, but the final recovery (36.0 ± 1.3%) was lower than in the
absence of ferric.
No ferrous was detected (data not shown): the Fe2+ produced in
reaction (1) was rapidly oxidized to Fe3+ by ozone (reaction (2)).
The decrease in Cu recovery with increasing ferric concentration
could be attributed to the salting-out effect [19] whereby Fe pre-
cipitates at high ferric concentration. This phenomenon has not
been observed in previous research on the application of ozone
and ferric to chalcopyrite leaching, where the ferric source and ele-
mental composition and mineral phase of samples differed, the
leaching time was short (2 h), and there were too few replicates
Fig. 2. Schematic diagram of ozone leaching system. to detect differences in Fe concentration [17,25].

335
J. Wang, F. Faraji and A. Ghahreman Journal of Industrial and Engineering Chemistry 104 (2021) 333–344

Fig. 3. Mean (±SD) (a) Cu and (b) Fe extraction versus leach time at three temperatures (leaching conditions: 10.1 ppm ozone, 10 g/L chalcopyrite, 0.5 M H2SO4).

Table 2
Mean (±SD) oxidation–reduction potential (mV) at three temperatures during chalcopyrite leaching (conditions: 10.1 ppm ozone, 10 g/L chalcopyrite, 0.5 M H2SO4).

Temperature (°C) Time (h)


1 3 5 8 24
25 510 ± 11 541 ± 7 530 ± 8 528 ± 10 558 ± 13
50 493 ± 8 535 ± 13 545 ± 14 565 ± 8 558 ± 4
85 539 ± 11 540 ± 10 545 ± 16 538 ± 8 647 ± 17

perhaps because dissolved ozone approached saturation above


10 ppm.

Solid/liquid ratio
Both Cu and Fe extraction were highest after the 24-h leach at
5 g/L S/L ratio, and Cu recovery was higher than Fe recovery
(Fig. 6). A moderate chalcopyrite concentration appears to be opti-
mum for mass transfer, and the unit quantity of ozone is greater
[26,32]. A 5 g/L S/L ratio was used in subsequent ozone leaching
tests.

Sulfuric acid concentration


The optimum H2SO4 concentration for Cu and Fe extraction was
0.5 M, and the lowest recoveries were observed at 2 M H2SO4
(Fig. 7). At high acid concentration, regeneration of H2SO4 may
cause saturation of SO2– +
4 and free H in the medium, which can
slow the leaching reactions and decrease the availability of ozone
as the main leaching agent. Moreover, it is possible for iron to pre-
Fig. 4. Mean (±SD) Cu extraction versus leach time at no ferric addition and three cipitate with SO2–4 to form cuprous and iron sulfate [26]. This
other ferric concentrations (leaching conditions: 25 °C, 10.1 ppm ozone, 10 g/L decrease in dissolved ozone solubility with increasing acid concen-
chalcopyrite, 0.5 M H2SO4). tration would need to be considered in the optimization [19].

Leaching time
Ozone concentration Under optimal leaching conditions, both Cu and Fe extraction
Both Cu and Fe extraction increased with increasing ozone con- increased in a near-linear fashion during a 48-h leach, reaching
centration (Fig. 5), which is expected since more oxidizing condi- nearly 100% recovery for Cu and 76.2 ± 1.7% recovery for Fe
tions enable better metal recovery from sulfide ores [31]. Overall, (Fig. 8). Nearly complete stoichiometric Cu extraction from CuFeS2
Cu recovery (Fig. 5 (a)) was higher than Fe recovery (Fig. 5 (b)). in mildly acidic conditions at room temperature has never been
The Cu recovery after 24 h at 2.2 ppm ozone was similar to that reported [33–35]. The residue was FeS2 (section 3.4), which we
after 1 h at 19.3 ppm ozone, consistent with previous reports that attribute to the galvanic interaction between chalcopyrite and pyr-
the mass transfer is greater at higher ozone concentrations, making ite [2,36].
the leach more intense and leading to a shorter treatment time On the basis of the leaching trials above, the optimum condi-
[18,23]. After 24 h, the Cu recovery was only slightly higher in tions for maximum Cu recovery that balanced economic consider-
the 19.3 ppm ozone treatment than the 10.1 ppm ozone treatment, ations were: leaching temperature of 25 °C; no addition of ferric
336
J. Wang, F. Faraji and A. Ghahreman Journal of Industrial and Engineering Chemistry 104 (2021) 333–344

Fig. 5. Mean (±SD) (a) Cu and (b) Fe extraction versus leach time at four ozone concentrations (leaching conditions: 25 °C, no added ferric, 10 g/L chalcopyrite, 0.5 M H2SO4).

Fig. 6. Mean (±SD) (a) Cu and (b) Fe extraction versus leaching time at three solid/liquid ratios (leaching conditions: 25 °C, no added ferric, 10.1 ppm ozone, and 0.5 M H2SO4).

Fig. 7. Mean (±SD) (a) Cu and (b) Fe extraction versus leaching time at four sulfuric acid concentrations (leaching conditions: 25 °C, no added ferric, 10.1 ppm ozone, and 5 g/L
chalcopyrite).

337
J. Wang, F. Faraji and A. Ghahreman Journal of Industrial and Engineering Chemistry 104 (2021) 333–344

Fig. 8. Mean (±SD) Cu and Fe extraction versus leaching time under optimal
leaching conditions: 25 °C, no added ferric, 10.1 ppm ozone, 5 g/L chalcopyrite, and Fig. 10. Sulfur removal from chalcopyrite in three leaching systems at three
0.5 M H2SO4. temperatures.

sulfate; 10.1 ppm ozone; 5 g/L S/L ratio; 0.5 M H2SO4 concentra- tion. Therefore, there is practically no iron recovery from either
tion; and 48 h leach time. of the systems as most of them precipitated out. The amount of
precipitate decreased with increasing temperature, and
Leaching of chalcopyrite in different oxidized systems 3.3 ± 0.7% iron extraction (only one) was attained at 85 °C in the
presence of ozone. Comparing Fig. 9 (a) with Fig. 3 (b), it is con-
Due to the presence of oxygen and regeneration of ferric in cluded that Fe leaching from chalcopyrite is impeded by the addi-
ozone–oxygen test (Fig. 3) and their ability to leach chalcopyrite tion of ferric sulfate. Comparing all the leaching recoveries of Cu
(Eqs. (3) and (4)), the roles of individual oxidants of ozone, oxy- and Fe in the three leaching systems at three temperatures,
gen, and ferric and their interactions remained unclear. To clear Table S1 shows that ozone-oxygen system at 85 °C presents opti-
this out, some tests were performed to compare two leaching mum condition (Cu and Fe extraction at 56.5% and 62.7%, respec-
systems of ozone-oxygen-ferric and oxygen-ferric at different tively). However, the system operated at 25 °C showed 45.0% and
temperatures using the most inhibitory ferric concentration 36.2% of Cu and Fe extractions, respectively, lower than those
(10 g/L; Fig. 4). obtained at 85 °C but the selectivity of Cu over Fe is significantly
Copper extraction in ozone-oxygen-ferric (Fig. 9 (a)) and higher. In addition, leaching at 25 °C can be less energy intensive
oxygen-ferric (Fig. 9 (b)) leaching systems were both temperature- and more eco-friendly than 85 °C, and the ozone concentration
dependent. The initial increase in extraction was similar in both was more stable without decomposition. Hence, ozone leaching
systems, but the highest recovery (55.6 ± 2.2%) was achieved with- of chalcopyrite at room temperature has been demonstrated as
out ozone at 85 °C, probably due to the higher stability of oxygen an environmentally attractive and technically feasible process with
than ozone at higher temperatures. On the other hand, Fe precipi- high potential for future studies.
tation occurs at all temperatures in both ozone–oxygen–ferric and At each of the three temperatures tested, sulfur extraction
oxygen–ferric leaching processes due to the ferric loss from solu- was approximately 3–4 times higher for tests with ozone than

Fig. 9. Mean (±SD) Cu extraction versus leach time at three temperatures using 10 g/L chalcopyrite, 0.5 M H2SO4, 1 L/min oxygen, and 10 g/L ferric (a) with and b) without
10.1 ppm ozone.

338
J. Wang, F. Faraji and A. Ghahreman Journal of Industrial and Engineering Chemistry 104 (2021) 333–344

Table 3
Mass of sulfur species in residues after leaching with 10.1 ppm ozone and four ferric concentrations at 25 °C.

Ferric concentration (g/L) Residue (g) Total S (ST, g) Elemental S (S0, g) S removal (%)
No ferric 5.1384 1.8823 0.0597 45.1
1 5.7985 2.3348 0.0795 31.9
5 6.2292 2.6813 0.1307 21.8
10 6.7542 2.7708 0.1898 19.2

Fig. 11. Shrinking core model for diffusion through the product layer (left column) surface chemical reaction (right column) control of copper leaching at three reaction
temperatures for (a) ozone-oxygen; (b) oxygen-ferric; (c) ozone-oxygen-ferric systems.

339
J. Wang, F. Faraji and A. Ghahreman Journal of Industrial and Engineering Chemistry 104 (2021) 333–344

tests without ozone (Fig. 10), indicating that ozone has the
highest oxidation potential among the three oxidizing agents
[23]. The inhibitory effect of ferric was evident only at the
two lower temperatures, whereas at 85 °C, ferric enhanced sul-
fur removal. The highest sulfur removal overall was seen at
85 °C in the presence of all three oxidants and is attributed
to their synergistic effects. Sulfidewas oxidized by ferric and
oxygen to elemental sulfur (Eqs. (3) and (4)), which formed a
coating on the chalcopyrite surface that limited further leaching.
The elemental sulfur was further oxidized to sulfuric acid by
ozone (Eq. (6)). This was confirmed by the initial pH increase
followed by a gradual decrease. The variation of ORP value dur-
ing chalcopyrite leaching in the three oxidizing systems at 25 °C
is shown in Figure S1.
The residue analysis also showed that the maximum sulfur
removal was achieved in the absence of ferric (Table 3) and
decreased with increasing ferric concentration, similar to the Cu
extraction trend during ozone leaching (reaction (1)). Elemental
sulfur did not limit ozone leaching, which agrees with the previous
research [11]. The mass of elemental sulfur increased slightly with
ferric concentration, confirming that ferric accumulation/precipita-
tion blocks or slows ozone diffusion.

Kinetic analysis

Shrinking core model is the most common model for evaluation


of kinetics data in the particle scale level. In particle level model,
reagents are required to diffuse (through liquid/solid) and reach
to the unreacted core of the particles and participate in the chem-
ical reaction. Leaching of Cu from chalcopyrite is a solid–liquid
heterogeneous reaction, and the leaching rate can be controlled
by either diffusion through the product layer (Eq. (8)) or surface
chemical reactions (Eq. (9)) according to the shrinking core model
[37,38]:
2
kP t ¼ 1  3ð1  X Þ3 þ 2ð1  XÞ ð8Þ
1
kR t ¼ 1  ð1  XÞ3 ð9Þ
where kP and kR are the apparent rate constants (slope of regres-
sion line) for the product layer and chemical reaction controlling
steps, respectively, and t is the time to achieve X fraction metal
recovery. The reaction rate constant (kd) depends on temperature
following the Arrhenius equation:
 
Ea
kd ¼ Aexp ð10Þ
RT
where A is the frequency factor, Ea is the apparent activation
energy of the reaction, R is the universal gas constant, and T is
the absolute temperature.
Both control models (Eqs. (8) and (9)) closely fit the empirical
data from Fig. 3 (a) and Fig. 9, indicating that all leaching systems
were under mixed control (Fig. 11). For each temperature, the
highest rate constants were associated with the ozone-oxygen
mixture (Fig. 11 (a)), which demonstrates the fastest Cu recovery.
Ozone diffusion was inhibited by the presence of ferric, especially
at 25 °C (Fig. 11 (a) and (c)).
From the Arrhenius equation (Eq. (10)), the ozone-oxygen
leaching activation energy (Ea) was the lowest (Fig. 12). Inhibition Fig. 12. Arrhenius plots for leaching of Cu in diffusion control model: (a) ozone-
by ferric sulfate at 25 °C mainly focuses on ozone diffusion, and oxygen; (b) oxygen-ferric; (c) ozone-oxygen-ferric.

340
J. Wang, F. Faraji and A. Ghahreman Journal of Industrial and Engineering Chemistry 104 (2021) 333–344

Fig. 13. X-ray diffraction patterns of chalcopyrite-leached residues: (a) 25 °C ozone-oxygen; (b) 25 °C ozone-oxygen-ferric; (c) 85 °C ozone-oxygen; and (d) 85 °C ozone-
oxygen-ferric.

Fig. 14. Scanning electron micrographs of chalcopyrite-leached residues: (a) 25 °C ozone-oxygen; (b) 25 °C ozone-oxygen-ferric; (c) 85 °C ozone-oxygen; and (d) 85 °C ozone-
oxygen-ferric.
341
J. Wang, F. Faraji and A. Ghahreman Journal of Industrial and Engineering Chemistry 104 (2021) 333–344

elevated temperature is chemically controlled due to ozone


decomposition.
Kinetic analysis of systems with redox couples (e.g., Fe3+/Fe2+ in
this study) must consider oxidation and reduction of the redox
couple along with dissolution of minerals [39]. Fe3+ reacts with
the mineral and mobilizes Cu. In the process, Fe3+ is reduced to
Fe2+, which is then oxidized by ozone back to Fe3+. Nicol [40] sug-
gested three scenarios for this dynamic interaction.

1. Type 1: oxidation of mineral and redox couple are the main


reactions and reduction of redox couple does not influence
the leaching rate;
2. Type 2: all the three reactions occur at a similar rate and affect
the overall leaching rate;
3. Type 3: half reactions associated with oxidation of the redox
couple are more significant than mineral dissolution.
The leaching systems presented here followed the Type 1 sce-
nario: all mobilized iron was measured as Fe3+, indicating the pres-
ence of enough oxidizing agent in the medium and rapid
conversion of the redox couple.

Characterization of leached residues

The chalcopyrite was partly dissolved in all four treatments, as


evidenced by the low CuFeS2 to FeS2 ratio in the residues (Fig. 13).
The relative intensity of the chalcopyrite and pyrite main peaks
was lower in 85 °C than 25 °C treatments, which indicates that
the higher temperature enhanced Cu leaching. The main phase
species were not affected at 25 °C (Fig. 13 (a) and (b)), whereas ele-
mental sulfur was only found in the 85 °C leached residues (Fig. 13
(c) and (d)). Since oxidation of sulfide by ozone can produce sulfu-
ric acid (Eq. (6)), the presence of elemental sulfur indicates less
ozone was available at 85 °C because it decomposes at high tem-
perature, as discussed in section 3.1.1. At 25 °C, the higher chal-
copyrite content of leached residue by ozone-oxygen-ferric than
ozone-oxygen can be explained by the corresponding lower Cu
recovery and no Fe removal. At 85 °C, the similar content of chal-
copyrite concurs with the corresponding Cu recovery; the higher
content of elemental sulfur in leached residue by ozone-oxygen-
ferric than ozone-oxygen was probably attributed to the higher S
removal (Fig. 10) and ferric involvement (Eq. (4)).
Ozone leaching at 25 °C resulted in a smooth surface on the
chalcopyrite with some microcracks (Fig. 14 (a) and (b)). The high
concentration of Fe in (b) relative to (a) based on energy dispersive
X-ray spectroscopy mapping (29.5 vs 29.0%) indicates some ferric
precipitated on the mineral surface and suppressed ozone diffusion
into the interior, which delayed the leaching in accordance with
the diffusion control mechanism. Fig. 14 (c) and (d) show extensive
small globules of sulfur on the surface of 85 °C leached residues,
which concurs with the XRD results and the proposed mechanism.
The very porous structure produced at 85 °C was caused by mas-
sive dissolution of chalcopyrite.
Fig. 15. Particle size distribution (PSDs) of the leached residues by (a) ozone-
At 25 °C without ferric and relative to the original chalcopyrite
oxygen at 25 °C; (b) ozone-oxygen-ferric at 25 °C; and (c) ozone-oxygen-ferric at (d80 = 60 mm, specific surface area 653.3 m2/kg), the particle size of
85 °C. ozone-leached residue was small (d80 = 44.9 mm) and the specific

342
J. Wang, F. Faraji and A. Ghahreman Journal of Industrial and Engineering Chemistry 104 (2021) 333–344

Fig. 16. Schematic of chalcopyrite leaching mechanisms under various oxidative conditions.

Table 4
Green chemistry metrics for copper extraction from chalcopyrite under optimal ozone leaching conditions*.

Material Yield (%) Conversion (%) Selectivity RME (%) AE (%) OE (%) PMI
CuFeS2 100 100 1.00 8.65 17.56 49.26 3.28

* Leaching temperature 25 °C; no addition of ferric sulfate; 10.1 ppm ozone; 5 g/L chalcopyrite; 0.5 M H2SO4 concentration; and 48 h leach time.

surface area was low (574.2 m2/kg; Fig. 15 (a)) due to the removal measure the in situ dissolved ozone concentration. In this mixed
of Cu (45%), Fe (38%), and S (45%). However, the addition of 10 g/L mechanism, the elemental sulfur layer is the main diffusion obsta-
ferric at 25 °C only caused a slight decrease in particle size (d80 = cle, and the reactions responsible for chalcopyrite oxidation are the
57 mm) and specific surface area (604.1 m2/kg; Fig. 15 (b)), indica- slowest. The oxygen-ferric leaching system has high copper extrac-
tive of iron precipitation with 18% Cu and 38% S extraction. Under tion at high temperature, but elemental sulfur is still a limiting
the same conditions at 85 °C (Fig. 15 (c)), the particle size showed factor.
little reduction (d80 = 59.3 lm), but the specific surface area was
very low (538.9 m2/kg). The high Cu (51%) and S (58%) removal Green chemistry metrics
made the surface very porous. All together, the SEM, XRD, and
PSD results also support the kinetics data, and indicate that both Green chemistry metrics for the optimum leaching conditions
diffusion and chemical reactions are playing roles in the leaching are summarized in Table 4. Detailed calculations can be found in
of copper. supporting information (Table S2 to Table S4 and Eq. S1 to Eq.
S8). The percentage yield, conversion, and selectivity values indi-
cate that this process is highly efficient in terms of molar and mass
Proposed leaching mechanism balance of the limiting step. The reaction mass efficiency (RME)
and atom economy (AE) are calculated based on the mass and
Dissolution of copper from chalcopyrite by ozone is a simple number of moles of the reagents and product, respectively. The
electrochemical leaching process, but more reactions at elevated RME to AE ratio expressed as a percent represents the optimum
temperature and (or) the presence of ferric complicate the process efficiency (OE). The process mass intensity (PMI) indicates that a
in various ways depending on leaching conditions (Fig. 16). Ozone relatively volume of other reagents (e.g., solvent, catalyst) are
plays a key role in improving the recovery of Cu by eliminating the required for complete copper dissolution from chalcopyrite. Over-
passive elemental sulfur layer that forms when sulfide is oxidized all, the green metrics show that copper leaching from chalcopyrite
to sulfuric acid. However, high temperature weakens the role of under optimal conditions is an economic, energy-efficient, and sus-
ozone because of significant decrease in ozone solubility [41], tainable process that has potential for future application (Fig. 16).
thereby changing the leaching mechanism to oxygen and ferric
oxidative leaching. The presence of ferric attenuates ozone oxida-
Conclusions
tion of chalcopyrite because ferric precipitation hinders ozone
diffusion.
(1) The inhibitory effect of ferric sulfate on chalcopyrite leach-
The kinetics analysis demonstrated that the leaching systems
ing was attributed to low ozone availability (diffusion con-
follow the shrinking core model with both diffusion control and
trol mechanism), excess SO2– 4 ions, and (or) the salting-out
chemical control. However, it is impossible to quantitatively ana-
effect of Fe3+.
lyze the mechanism since there is no efficient way to continuously
343
J. Wang, F. Faraji and A. Ghahreman Journal of Industrial and Engineering Chemistry 104 (2021) 333–344

(2) Among all process variables tested, temperature had the [9] M.A. Ghomi, M. Mozammel, H. Moghanni, L. Shahkar, Hydrometallurgy 189
(2019), https://doi.org/10.1016/j.hydromet.2019.105120.
greatest effect on copper extraction and determined the rel-
[10] G. Nazari, D.G. Dixon, D.B. Dreisinger, Hydrometallurgy 105 (2011) 251–258,
ative effectiveness of the three oxidants tested. https://doi.org/10.1016/j.hydromet.2010.10.013.
(3) All leaching processes followed the shrinking core model of [11] T. Havlik, M. Skrobian, Can. Metall. Q. 29 (1990) 133–139, https://doi.org/
mixed diffusion control and chemical reaction control. The 10.1179/cmq.1990.29.2.133.
[12] M. Al-Harahsheh, S. Kingman, A. Al-Harahsheh, Hydrometallurgy 91 (2008)
overall reaction rate at 25 °C was controlled by ozone diffu- 89–97, https://doi.org/10.1016/j.hydromet.2007.11.011.
sion through solids, whereas at 85 °C, it was controlled by [13] M. Sokić, B. Marković, S. Stanković, Z. Kamberović, N. Štrbac, V. Manojlović, N.
oxygen and ferric availability as the dominant oxidants. Petronijević, Metals 9 (2019) 1–13, https://doi.org/10.3390/met9111173.
[14] T. Havlik, M. Laubertova, A. Miskufova, J. Kondas, F. Vranka, Hydrometallurgy
Characterization of the leach residues confirmed the leach- 77 (2005) 51–59, https://doi.org/10.1016/j.hydromet.2004.10.009.
ing mechanisms. [15] L. Chang-Li, X. Jin-Lan, N. Zhen-Yuan, Y. Yi, M. Chen-Yan, Bioresour. Technol.
(4) Among three oxidizing agents, ozone exhibited the strongest 110 (2012) 462–467, https://doi.org/10.1016/j.biortech.2012.01.084.
[16] S. Rakovsky, M. Anachkov, G. Zaikov, Chem. Chem. Technol. 3 (2009) 139–161,
oxidizing ability in terms of converting sulfur to sulfate ions. http://ena.lp.edu.ua:8080/handle/ntb/1316.
Oxygen and ferric contributed to the formation of elemental [17] F.R.C. Pedroza, M.D.J.S.; Aguilar, T.P. Treviño, A.M. Luévanos, M.S. Castillo, Min.
sulfur. Proc. Ext. Met. Rev. 33 (2012) 269–279. https://doi.org/10.1080/
08827508.2011.584093.
(5) Ozone leaching at low temperature (25 °C) offered the best [18] F. Nava, A. Uribe, R. Perez, Eur. J. Miner. Process. Environ. Prot. 3 (2003) 316–
trade-off between copper recovery and economic and envi- 323, http://www.absoluteozone.com/assets/ozone_cyanide_removal_july_29.
ronmental considerations as demonstrated by green chem- pdf.
[19] M.Z. Mubarok, K. Sukamto, Z.T. Ichlas, A.T. Sugiarto, Miner. Metall. Proc. 35
istry metrics for the optimum conditions: 100% copper
(2018) 133–140, https://doi.org/10.19150/mmp.8462.
extraction was achieved after 48 h leaching with no addition [20] M. Ukasik, T. Havlik, Hydrometallurgy 77 (2005) 139–145, https://doi.org/
of ferric, 10.1 ppm ozone in the ozone-oxygen mixture, 5 g/L 10.1016/j.hydromet.2004.10.017.
chalcopyrite, and 0.5 M H2SO4. [21] Q. Tian, H. Wang, Y. Xin, D. Li, X. Guo, Hydrometallurgy 159 (2016) 126–131,
https://doi.org/10.1016/j.hydromet.2015.11.011.
(6) The optimized ozone leaching process was highly efficient, [22] C. Rodríguez-Rodríguez, F. Nava-Alonso, A. Uribe-Salas, Hydrometallurgy 149
and further work is required to determine if scale-up will (2014) 168–176, https://doi.org/10.1016/j.hydromet.2014.08.006.
make it feasible for future applications. [23] Q. Li, D. Li, F. Qian, Hydrometallurgy 97 (2009) 61–66, https://doi.org/10.1016/
j.hydromet.2009.01.002.
[24] C. Rodríguez-Rodríguez, F. Nava-Alonso, A. Uribe-Salas, Can. Metall. Q. 57
(2018) 294–303, https://doi.org/10.1080/00084433.2018.1460437.
Declaration of Competing Interest [25] F.R. Carrillo-Pedroza, M.A. Sánchez-Castillo, M.J. Soria-Aguilar, A. Martínez-
Luévanos, E.C. Gutiérrez, Can. Metall. Q. 49 (2010) 219–226, https://doi.org/
10.1179/cmq.2010.49.3.219.
The authors declare that they have no known competing finan- [26] T. Havlik, J. Dvorscikova, Z. Ivanova, R. Kammel, Metall 53 (1999) 57–60,
cial interests or personal relationships that could have appeared http://pascal-francis.inist.fr/vibad/index.php?
action=getRecordDetail&idt=1702049.
to influence the work reported in this paper. [27] L.N. Krylova, Russ. J. Non-Ferr Met. 61 (2020) 49–56, https://doi.org/10.3103/
S1067821220010095.
Acknowledgements [28] N. Hiroyoshi, S. Kuroiwa, H. Miki, M. Tsunekawa, T. Hirajima, Hydrometallurgy
74 (2004) 103–116, https://doi.org/10.1016/j.hydromet.2004.01.003.
[29] H. Naderi, M. Abdollahy, N. Mostoufi, M.J. Koleini, S.A. Shojaosadati, Z. Manafi,
The authors would like to acknowledge study funding from Int. J. Miner. Metall. 18 (2011) 638–645, https://doi.org/10.1007/s12613-011-
Hatch Ltd. (Mississauga, ON, Canada) and Mitacs through fund # 0489-7.
IT13551. [30] J. Wang, F. Faraji, A. Ghahreman, Minerals 10 (2020) 633, https://doi.org/
10.3390/min10070633.
[31] P.C. Hernández, J. Dupont, O.O. Herreros, Y.P. Jimenez, C.M. Torres, Minerals 9
Appendix A. Supplementary data (2019) 1–13, https://doi.org/10.3390/MIN9040250.
[32] X.Y. Guo, Y.T. Xin, W. Hao, Q.H. Tian, T. Nonferr, Metal. Soc. 27 (2017) 1888–
1895, https://doi.org/10.1016/S1003-6326(17)60213-9.
Supplementary data to this article can be found online at [33] D. Shin, J. Ahn, J. Lee, Hydrometallurgy 183 (2019) 71–78, https://doi.org/
https://doi.org/10.1016/j.jiec.2021.08.036. 10.1016/j.hydromet.2018.10.021.
[34] S. Zhong, Y. Li, J. Mater. Res. Technol. 8 (2019) 3487–3494, https://doi.org/
10.1016/j.jmrt.2019.06.020.
References [35] N.T.P. Thao, S. Tsuji, S. Jeon, I. Park, C.B. Tabelin, M. Ito, N. Hiroyoshi,
Hydrometallurgy 194 (2020), https://doi.org/10.1016/j.hydromet.2020.
[1] S. Wang, JOM 57 (2005) 48–51, https://doi.org/10.1007/s11837-005-0252-5. 105299.
[2] S.M.J. Koleini, V. Aghazadeh, A. Sandström, Miner. Eng. 24 (2011) 381–386, [36] N.V. Fomchenko, M.I. Muravyov, Hydrometallurgy 185 (2019) 82–87, https://
https://doi.org/10.1016/j.mineng.2010.11.008. doi.org/10.1016/j.hydromet.2019.02.002.
[3] Y. Li, N. Kawashima, J. Li, A.P. Chandra, A.R. Gerson, Adv. Colloid Interface Sci. [37] F. Faraji, A. Alizadeh, F. Rashchi, N. Mostoufi, Rev. Chem. Eng. (in press)
197 (2013) 1–32, https://doi.org/10.1016/j.cis.2013.03.004. https://doi.org/10.1515/revce-2019-0073.
[4] A. Ghahremaninezhad, D.G. Dixon, E. Asselin, Electrochim. Acta 87 (2013) 97– [38] O. Levenspiel, Chemical Reaction Engineering, third ed., John Wiley & Sons,
112, https://doi.org/10.1016/j.electacta.2012.07.119. New York, 1999.
[5] Y.J. Xian, S.M. Wen, J.S. Deng, J. Liu, Q. Nie, Can. Metall. Q. 51 (2012) 133–140, [39] A. Ghahremaninezhad, D.G. Dixon, E. Asselin, Hydrometallurgy 125 (2012) 42–
https://doi.org/10.1179/1879139512Y.0000000001. 49, https://doi.org/10.1016/j.hydromet.2012.05.004.
[6] M.D. Sokić, B. Marković, D. Živković, Hydrometallurgy 95 (2009) 273–279, [40] M.J. Nicol, The Role of Electrochemistry in Hydrometallurgy. In 4th
https://doi.org/10.1016/j.hydromet.2008.06.012. International Symposium on Hydrometallurgy; AIME: Salt Lake City, Utah,
[7] H.S. Yoon, C.J. Kim, K.W. Chung, J.Y. Lee, S.M. Shin, S.R. Kim, M.H. Jang, J.H. Kim, USA, 1993.
S. Lee, S.J. Yoo, Korean J. Chem. Eng. 34 (2017) 1748–1755, https://doi.org/ [41] A.K. Biń, Ozone Sci. Eng. 28 (2006) 67–75, https://doi.org/10.1080/
10.1007/s11814-017-0053-x. 01919510600558635.
[8] T. Wen, Y. Zhao, Q. Xiao, Q. Ma, S. Kang, H. Li, S. Song, Results Phys. 7 (2017)
2594–2600, https://doi.org/10.1016/j.rinp.2017.07.035.

344

You might also like