You are on page 1of 8

Hydrometallurgy 183 (2019) 71–78

Contents lists available at ScienceDirect

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

Kinetic study of copper leaching from chalcopyrite concentrate in alkaline T


glycine solution

Doyun Shina,b, Junmo Ahna, Jaeheon Leea,
a
Department of Mining and Geological Engineering, University of Arizona, Tucson, USA
b
Mineral Resources Research Division, Korea Institute of Geoscience and Mineral Resources (KIGAM), Daejeon, Republic of Korea

A R T I C LE I N FO A B S T R A C T

Keywords: Recently, glycine, one of the simplest amino acids, has been emerging as an efficient and environmental-friendly
Chalcopyrite lixiviant for copper leaching. In the present study, an alkaline glycine–peroxide lixiviant system was used to
Glycine leach copper from chalcopyrite concentrate by a suite of kinetic studies under different conditions. The particle
Alternative lixiviant size of the concentrate was P80 40 μm. Glycine concentration was varied from 0.5 to 3.0 M (glycine to copper
Passivation
molar ratio of 1.1 to 6.6:1) and the pH was maintained at 11 during the experiment. The copper extraction
within 96 h was observed from 13 to 14% with glycine concentrations from 0.5 to 2 M, while the copper ex-
traction decreased at 3 M of glycine. By reconstituting the glycine concentration during the leach test and adding
hydrogen peroxide periodically, the copper extraction with 1 M glycine at 96 h increased from 14 to 21%,
respectively. By changing the glycine and hydrogen peroxide solution every 24 h, 42% of copper extraction was
achieved within 168 h. With increasing temperature, copper extraction decreased, while iron dissolution in-
creased. Copper sulfide and copper sulfate precipitates were found in the leaching liquor at 22 °C test. Poorly
crystalline copper-iron sulfate phase was observed in the leaching liquor at 35 and 45 °C. This precipitation is
believed to be one of the factors to decrease the copper recovery from chalcopyrite.

1. Introduction this passivation to increase the leaching efficiency. To overcome this


passivation, it is required to artificially control redox potential during
The greatest global resource of copper is low-grade chalcopyrite and the bioleaching process (Yu et al., 2011) or use thermo-acidophilic
heap bioleaching may be the only economic option for its processing. system (Gericke et al., 2009) to keep the redox potential low and the
However, formation of a passivating layer on the surface of chalcopyrite ferrous ion concentration high. Addition of silver and chloride ions as
slows down the oxidation reaction and reduces the efficiency of the catalyst to increase the solubility of metal was reported (Feng et al.,
process. 2013) resulting in less production of S0 membrane and jarosite pre-
cipitation. Panda et al. proposed to wash the system with sodium and
CuFeS2 + 4 Fe3 + → Cu2 + + 5 Fe2 + + 2S° (1)
ammonium carbonate to remove the passivation layer (Panda et al.,
Passivation phenomenon is traditionally described in ferric medium 2013).
following Eq. (1). There is a diffusion barrier of polysulfide or elemental Recently, glycine, one of the simplest amino acids, has been emer-
sulfur between the leaching solution and the chalcopyrite that slows ging as an efficient and environmentally friendly lixiviant to leach
down the dissolution rate. Iron deposition on the surface of the ore is copper (Oraby and Eksteen, 2014; Tanda et al., 2017). Comparing with
also a very important issue in passivation. It is clear that chalcopyrite other amino acids, glycine is relatively inexpensive and the perfor-
passivation is related to high redox potentials or high Fe3+/Fe2+ ratios mance was proved (Eksteen and Oraby, 2015). In alkaline condition
(Córdoba et al., 2008). If the initial potential is very high, equilibrium (above pH 10), most of glycine exists as an anion (H2NCH2COO−), and
of ferric/ferrous sulfate moves and rapid precipitation of ferric ion as forms a strong complex with copper (I) and (II) (Eqs. (2)–(4)) (Aksu and
jarosite occurs resulting in the formation of passivation layer. At high Doyle, 2001).
solution potential (i.e., > 600 mV vs. Ag/AgCl), a passivation layer
Cu2 + + H2 NCH2 COO− → Cu(NH2 CH2 COO)+ (2)
could be formed.
Many alternative leaching systems have been investigated to avoid Cu2 + + 2H2 NCH2 COO− → Cu(NH2 CH2 COO)2 (3)


Corresponding author.
E-mail address: jaeheon@email.arizona.edu (J. Lee).

https://doi.org/10.1016/j.hydromet.2018.10.021
Received 26 January 2018; Received in revised form 18 October 2018; Accepted 26 October 2018
Available online 29 October 2018
0304-386X/ © 2018 Elsevier B.V. All rights reserved.
D. Shin et al. Hydrometallurgy 183 (2019) 71–78

Table 1
Elemental analysis of chalcopyrite concentrate used in this study.
Element (%)

Cu S Fe Al Ca K Mg Na Ag (ppm) As (ppm) Pb (ppm) Zn (ppm) Mo (ppm)

Content 28.8 31.1 26.4 0.62 0.28 0.38 0.09 0.14 90.5 194 582 1780 1890

25 25
(a) 0.5 M glycine (Figure 2)
10 g/L H2SO4 and 10 g/L ferric
Constant 0.5 M glycine
1 M glycine and 3% H2O2
1 M glycine (Figure 2)
20 20
Constant 1 M glycine

Cu extracted (%)
Cu extracted (%)

15 15

10 10

5 5

0 0
0 20 40 60 80 100 0 20 40 60 80 100

Leaching time (hr) Leaching time (hr)

Fig. 1. Copper extraction by H2SO4-ferric and glycine-peroxide from the chal- 25


copyrite concentrate: 6% pulp density, 10 g/L H2SO4 + 10 g/L ferric or 1 M (b)
glycine +3% hydrogen peroxide, room temperature, bottle roll.
20
25
Cu extracted (%)

0.5 M glycine and 3% H2O2


0.5 M glycine and 5% H2O2 15
20 1 M glycine and 3% H2O2
Cu leaching efficiency (%)

1 M glycine and 5% H2O2


2 M glycine and 3% H2O2 10
15 3 M glycine and 3% H2O2
w/o glycine, 3% H2O2
5 0.5 M glycine, periodic addition of 3% H2O2
10 1 M glycine, periodic addition of 3% H2O2
Constant 1 M Glycine, periodic addition of 3% H2O2
0
5 0 20 40 60 80 100

Leaching time (hr)

0 Fig. 3. Effect of (a) glycine or (b) hydrogen peroxide addition during the
0 20 40 60 80 100 leaching test on copper extraction: 0.5, 1, 2, and 3 M glycine, 3% H2O2 (a) or
1% H2O2 was added at 0, 6, and 24 h (total 3%, b), pH 11, room temperature,
Leaching time (hr)
bottle roll.
Fig. 2. Effect of glycine and hydrogen peroxide concentration on copper ex-
traction: 0.5 and 1 M glycine, 3 and 5% H2O2, pH 11, room temperature, bottle complex (Hariharaputhiran et al., 2000). Thus, the generation of these
roll.
hydroxyl radicals increases copper dissolution in glycine-hydrogen
peroxide system. Hydrogen peroxide is also environmentally benign
Cu+ + 2H2 NCH2 COO− → Cu(NH2 CH2 COO)2− (4) because it decomposes to water.
Very few studies have been reported on copper leaching from
This complexation reaction can be enhanced by addition of hy-
chalcopyrite in the alkaline glycine-peroxide system. Eksteen et al.
drogen peroxide as an oxidant. The oxidation rate of copper is increased
(2017) have attempted to recover copper from chalcopyrite and found
because the anodic reaction is inhibited and the cathodic reaction (Eqs.
that alkaline atmospheric pre-oxidation and ultra-fine grinding was
(5)–(6)) is enhanced in the presence of hydrogen peroxide (Du et al.,
needed to obtain higher leaching efficiency by 40 g/L glycine with
2004).
oxygenation (0.75 mL/min O2) at 60 °C.
H2 O2 + 2e− → 2OH− (5) In the present study, the authors investigated an alkaline glycine–-
peroxide system to leach copper from chalcopyrite by a suite of kinetic
H2 O2 ± e− → OH·± OH– (6)
studies under different conditions. The leaching solution, leached re-
Also, under alkaline conditions, hydroxyl radicals (OH·) which are sidue and byproducts were analyzed by various instrumental techni-
more powerful oxidizing agents than hydrogen peroxide can be pro- ques.
duced from hydrogen peroxide by catalysis of copper-amino acid

72
D. Shin et al. Hydrometallurgy 183 (2019) 71–78

40 400 50
Cu extracted - Constant 1 M glycine, 3% H2O2
Cu extracted - Constant 1 M glycine, periodic 3% H2O2
Glycine consumed - Constant 1 M glycine, 3% H2O2

Glycine consumption (mmol)


40

Cu extraction efficiency (%)


30 Glycine consumed - Constant 1 M glycine, periodic 3% H2O2 300
Cu extracted (mmol)

30
20 200

20

10 100

10
96 hr leach w/o changing lixiviants
0 0 Changing lixiviants every 24 hrs
0 20 40 60 80 100
0
Leaching time (hr) 0 20 40 60 80 100 120 140 160 180

Fig. 4. Comparision between the amount of copper extracted and glycine Leaching time (hr)
consumed during the leaching of Fig. 3 (a) and (b). Full and open symbols
Fig. 6. Effect of changing a lixiviant and an oxidant on copper extraction: 6%
represent the amount of copper extracted and glycine consumed, respectively.
pulp density, 1 M glycine, 3% H2O2, pH 11, room temperature bottle roll.

25
(a) Cu copper mine in Southern Arizona. The elemental composition of the
RT bottle roll concentrate was analyzed using acid digestion followed by solution
35 C analysis for different metals using inductively coupled plasma optical
20
45 C
emission spectrometry (ICP-OES) (Table 1). X-ray diffraction showed
that the concentrate consisted of 89.5% chalcopyrite, 5.0% pyrite, 0.2%
Cu extracted (%)

15 molybdenite, 0.7% quartz, 3.1% K-Feldspar, and 1.5% plagioclase. The


particle size of the sample used in the leaching experiments was P80
40 μm.
10

2.2. Leaching study

5
A bottle roll leaching system was used to investigate the effect of
glycine and hydrogen peroxide concentration at room temperature. The
slurry was agitated by rolling the bottle on a rack at 60 rpm. For tem-
0
perature-controlled experiments, a 1-liter volume glass reactor
0 20 40 60 80 100
equipped with an aerator was used. The slurry was mixed with a
Leaching time (hr)
magnetic stirrer at 400 rpm and the temperature was maintained at 35
and 45 °C. The pulp density was 6.25% (20 g solid and 300 mL solu-
7.0
tion). Glycine concentration was varied from 0.5 to 3.0 M (glycine to
(b) Fe
RT bottle roll copper molar ratio = 1.1 to 6.6:1) and the hydrogen peroxide con-
6.5
35 C centration was changed between 3 and 5%. Kinetic samples were col-
0.8
45 C lected at 1, 6, 24, 48, 72, and 96 h of leaching using a syringe-mem-
brane filter of 0.45 μm pore size. The solution pH was maintained at
Fe extracted (%)

11.0 by adding 50% NaOH solution at every sampling point. The


0.6
average pH drop was observed to be about 0.5 in 24 h. Glycine con-
centrations in the kinetic samples were measured based on formol ti-
0.4
tration method (Zoecklein et al., 1995) and reconstituted to the initial
concentration if needed. The filtrates were analyzed for copper and iron
by atomic absorption spectrometry (AAS; Perkin-Elmer 2380, Perkin
0.2 Elmer, MA, USA) with air and acetylene gas. Dissolved oxygen, Eh, and
pH were monitored during the leaching experiments. Sulfuric acid was
also used as a lixiviant for copper leaching at 10 and 20 g/L with 10 g/L
0.0 ferric ion to compare the extraction efficiency.
0 20 40 60 80 100 The mineralogical compositions of the sample before and after
Leaching time (hr) leaching were analyzed by X-ray diffraction (XRD), scanning electron
microscope (SEM) and energy-dispersive X-ray spectroscopy (EDS). The
Fig. 5. Effect of temperature on copper extraction: 1 M glycine, 3% H2O2, chemical composition of the leaching residue and the precipitates were
pH 11, room temperature bottle roll, 35, and 45 °C (stirred reactor). analyzed by inductively coupled plasma mass spectrometer (ICP-MS;
Perkin-Elmer Elan DRC-II ICP-MS, Perkin Elmer, MA, USA) followed by
2. Materials and methods 3-acid digestion (fluoboric-nitric-hydrochloric).

2.1. Sample preparation 3. Results and discussion

Chalcopyrite flotation concentrate was received from an operating Fig. 1 shows < 2% of copper extraction by sulfuric acid with ferric

73
D. Shin et al. Hydrometallurgy 183 (2019) 71–78

Fig. 7. SEM images of the leaching residue after 168 h with changing the lixiviant, Magnification 1000×. (a) BSE, (b) EDS images. The detailed investigations are
given in the text.

ion as an oxidant while glycine with hydrogen peroxide dissolved up to peroxide concentrations is shown in Fig. 2. The molar ratio of glycine to
15% of copper within 96 h. The formation of an elemental sulfur layer copper was 1.1 to 6.6:1 (0.5–3 M). Without glycine, in the presence of
or iron deposition on chalcopyrite surface inhibited the reaction hydrogen peroxide only, no copper extraction was observed as ex-
(Munoz et al., 1979) thus copper extraction did not increase after 6 h in pected. The copper extraction within 96 h was observed from 13 to 14%
sulfuric acid and ferric solution. The effect of glycine and hydrogen with glycine concentrations from 0.5 to 2 M, while the copper

74
D. Shin et al. Hydrometallurgy 183 (2019) 71–78

peroxide (20%). Thus, both of glycine and hydrogen peroxide addition


enhanced the copper extraction and its kinetics.
Glycine consumption was also compared as shown in Fig. 4. At
every 24 h, the glycine consumption of the sample with constant 1 M
glycine and 3% hydrogen peroxide (Fig. 3(b) inverse open triangle) was
about 0.08–0.09 M, and the sample with constant 1 M glycine and
periodic 3% hydrogen peroxide (Fig. 3(b) solid square) was about
0.2–0.3 M, showing about 10 and 30% glycine was consumed by adding
hydrogen peroxide once or periodically within 24 h, respectively. At
every sampling point, the glycine concentration was reconstituted to
1 M. Even though the copper extraction efficiencies at each time point
were almost similar with 1 M glycine and periodic addition of 3% hy-
drogen peroxide (Fig. 3(b) inverse open triangle) and constant 1 M
glycine and 3% hydrogen peroxide (Fig. 3(a) open triangle), the glycine
consumption in the test with periodic addition of 3% hydrogen per-
oxide was higher than the test with constant glycine concentration.
Since the hydroxyl radicals oxidize glycine (Berger et al., 1999) and the
production of hydroxyl radicals from hydrogen peroxide is promoted by
copper-glycine complex (Du et al., 2004), glycine may be excessively
consumed when hydrogen peroxide is freshly added. Also, as shown in
Eqs. (2)–(4), one mole of copper should consume one or two moles of
glycine, however, the glycine consumption was about 10 times than the
copper extracted, possibly due to glycine oxidation by hydroxyl radical
and formation of ligand complex with other metals.
With increasing temperature (from room temperature to 45 °C), the
copper extraction decreased (Fig. 5 (a)), while the iron extraction in-
creased (Fig. 5 (b)). To control temperature, a 1-l volume stirred reactor
equipped with an aerator was used. These data indicate that copper
extraction is possibly inhibited, and the inhibition might be promoted
by increasing temperature or glycine concentration > 1 M. For com-
parison, pure copper powder was leached by 1 M glycine and 3% hy-
drogen peroxide at room temperature and 45 °C. After 1 h, the copper
extraction was 7% and 11% at room temperature and 45 °C, respec-
tively. Previous studies on chalcopyrite leaching in acid ferric sulfate
medium also reported that the activation energy during chalcopyrite
leaching with ferric ion is quite high, and the copper recovery rate
increased with temperature (Munoz et al., 1979; Córdoba et al., 2008).
Thus, it is believed that the lower copper extraction at high temperature
in glycine peroxide system is not common, but only observed in the
chalcopyrite leaching. Eksteen et al. reported that copper extraction
was enhanced by increasing temperature from room temperature to
60 °C and it is believed due to the different experimental condition of
the tests (Eksteen et al., 2017). In this study, air was only used to supply
oxygen, but pure oxygen was used in their study. In the case of other
copper minerals, since the glycine-peroxide system dissolved up to 80%
of total copper from the gold-copper concentrate including chalcocite,
Fig. 8. The images of (a) the leaching liquor freshly taken and after 24 h, (b) the Cu-metal, cuprite, chalcopyrite, and bornite (Oraby and Eksteen, 2014),
precipitate collected at room temperature, and (c) the precipitate at 45 °C. the effect of temperature on copper leaching was not investigated.
A series of tests were carried out using fresh lixiviant and the oxi-
extraction decreased at 3 M of glycine. The effect of hydrogen peroxide dant (i.e., the 1 M glycine and 3% hydrogen peroxide solution) by
on leaching efficiency was not significant. It is worth to note that the changing them every 24 h at room temperature. Copper extraction was
results shown in Fig. 2 were obtained without reconstituting the glycine 42% within 168 h (Fig. 6). At 96 h, the copper extraction was 29%,
concentration during the tests. indicating increase of copper extraction efficiency comparing with the
Copper extractions were calculated while maintaining the glycine 96 h leach test without changing the solution (21% on Fig. 3(a)). This
concentration by reconstituting it when kinetic samples were collected. result indicates that the leaching residue still contains glycine soluble
(Fig. 3 (a)). Comparing with the results shown in Fig. 2, the copper copper. After 168 h of leaching, the residue was analyzed by ICP
extraction with 1 M glycine at 96 h increased from 14 to 21%. As shown showing 18.6% Cu, 36.8% Fe, and 23.01% S. XRD analysis was also
in Fig. 3 (b), periodic addition of hydrogen peroxide in 24 h enhanced performed to see the mineralogy of the residue showing 2% quartz
copper extraction, from 14 to 20% in the presence of 1 M glycine. When (SiO2), 8% K-feldspar (KAlSi3O8), 6% plagioclase (NaAlSi3O8 –
the glycine concentration was 0.5 M, the copper extraction did not in- CaAl2Si2O8), 6% pyrite (FeS2), 4% bornite (Cu5FeS4), 54% chalcopyrite
crease much by reconstituting glycine concentration or periodic addi- (CuFeS2), and 20% amorphous fraction. Comparing with the chalco-
tion of hydrogen peroxide. In the constant 1 M glycine sample with pyrite content and Cu grade of the original concentrate (95 and 28.8%,
periodic addition of hydrogen peroxide, the rate of copper extraction respectively), 54% of chalcopyrite and 18.6% of copper in the residue
increased, but the copper extraction at 96 h (21%) was about the same are acceptable numbers based on 42% of copper extraction. Fig. 7
with the 1 M glycine sample with periodic addition of hydrogen shows the BSE (Back Scattered Electron) SEM images and EDS mapping
images of the residue. The unreacted chalcopyrite grains were observed

75
D. Shin et al. Hydrometallurgy 183 (2019) 71–78

Table 2
The (a) elemental and (b) mineralogical composition of the precipitates collected at room temperature (RT), 35 °C, and 45 °C.
(a) Elemental composition (%)

Cu S Fe Al Ca K Mg Na Ag (ppm) As Pb (ppm) Zn (ppm) Mo (ppm)


(ppm)

RT 61.8 15.2 0.45 – – 0.08 – 3.4 89.1 14.7 – – –


35 °C 31.7 10.7 15.4 0.56 0.36 0.22 0.22 3.9 151.3 93.7 0.11 0.03 0.16
45 °C 33.7 10.1 16.9 0.69 0.57 0.20 0.46 2.8 96.7 100.4 0.47 0.02 0.06

(b) Mineralogy composition (%)

RT 35 °C 45 °C

Plagioclase (NaAlSi3O8 – CaAl2Si2O8) – – 5


Siderite (FeCO3) – – 3
Chalconatronite (Na2Cu(CO3)2•3(H2O)) 5 – –
Pyrite (FeS2) – – 3
Chalcopyrite (CuFeS2) – 3 7
Chalcocite Group (Cu2S) 58 36 35
Perkovaite (CaMg2(SO4)3) – 2 –
Butlerite (Fe3+(SO4)(OH)•2(H2O)) 3 1 –
Poitevinite ((Cu,Fe2+,Zn)SO4•(H2O)) 3 – –
Bonattite (CuSO4•3(H2O)) 2 4 –
Alpersite ((Mg,Cu)SO4•7H2O) – 2 –
Antlerite (Cu3(SO4)(OH)4) 1 – –
(NH4)3Cu2(SO4)3(OH) 3 2 –
Kobyashevite (Cu5(SO4)2(OH)6·4H2O) 5 – –
Natrite (Na2CO3) 5 5 –
Amorphous fraction 16 46 47
Total 100 100 100

as shown at spot 1, and the poorly-crystalline materials associated with by hydrogen peroxide and addition of sodium hydroxide to adjust pH.
iron and oxygen as shown at spot 2 might be iron (hydr)oxide, possibly Since the hydroxyl radicals oxidize glycine (Berger et al., 1999) and the
this is the principal component of the 20% amorphous fraction de- production of hydroxyl radicals from hydrogen peroxide is promoted by
termined by XRD. Eksteen et al. (2017) has also reported that amor- copper-glycine complex (Du et al., 2004), carbonate can be produced
phous iron oxy-hydroxide was detected. from the oxidation reaction and reacted with high sodium from adding
Interestingly, a dark-brown precipitate was observed in the leaching sodium hydroxide. Sodium carbonate can crystallize within the porous
liquor after only a few minutes of sampling even though it was filtrated building materials when soluble sodium compounds in an alkaline en-
with 0.45 μm pore filter (Fig. 8 (a)). The precipitates were harvested vironment. Several small grains (~1 μm) of plagioclase (NaAlSi3O8 –
from the leaching liquor after 96 h at room temperature, 35, and 45 °C CaAl2Si2O8) were observed in the precipitate at 35 °C, and several
tests shown on Fig. 5, filtered with P4 grade filter paper (particle re- grains of hornblende (Ca2(Mg,Fe,Al)5(Al,Si)8O22(OH)2) and other sili-
tention 4–8 μm), settled down the precipitates followed by decanting cate material were observed below the detection limits of XRD in the
the supernatant, and centrifuged at 3000 rpm for 30 mins. The mor- precipitate at 45 °C.
phological characteristics of the precipitates in the leaching liquor at Based on the data of the chemical analysis, XRD, and SEM-EDS/BSE
different temperature were different. The precipitate collected from imaging analysis it was found that some of leached copper, iron, and
room temperature test was flakier and easily broken-up (Fig. 8 (b)), but oxidized sulfur may be precipitated after leaching occurred and the
the precipitates at 35 and 45 °C were harder solid (Fig. 8 (c)). The precipitation pattern changed in different leaching temperatures. At
chemical and mineralogical compositions of precipitates were analyzed room temperature, copper can be precipitated with sulfate and sulfide,
by chemical analysis and XRD. As shown in Table 2 (a), the precipitate which complexed with the byproduct of leaching, sulfur. This copper
at room temperature consists of mostly copper and sulfur, while the sulfide precipitation in alkaline condition has been previously reported
precipitates at 35 and 45 °C contain significant amount of iron with in enargite leaching (Baláž et al., 2000; Curreli et al., 2009). In enargite
copper and sulfur. In the mineralogical data, most of crystalline copper alkaline leaching by Na2S and NaOH, arsenic was totally decomposed
and sulfur were found to be chalcocite (Cu2S) in all the precipitates. The and new sulfide product, covellite (CuS) or Cu1.5S, was produced. The
content of chalcocite decreased with increasing temperature, while the formation of covellite was also found in the copper‑gold concentrate
content of amorphous fraction significantly increased with increasing leaching including chalcocite, Cu-metal, cuprite, chalcopyrite, bornite,
temperature. Some copper sulfate minerals such as bonattite, antlerite, and covellite (Oraby and Eksteen, 2014). Copper as this secondary
and kobyashevite were detected in the precipitate at room temperature covellite is concentrated in the leaching residue, and not further lea-
and 35 °C. To characterize the poorly crystalline fractions, SEM-EDS/ ched. The low leachability of secondary covellite as an intermediate
BSE imaging analysis was performed (Fig. 9 and Table 3). The poorly product of chalcocite leaching has been studied (Duda and Bartecki,
crystalline material in the precipitate at room temperature mostly 1982; Cheng and Lawson, 1991), however, the secondary covellite in
consisted of copper sulfide and copper sulfate, based on the elemental this study is newly crystalized from the dissolved copper of chalco-
composition from the SEM-EDS analysis. Several grains of chalcona- pyrite, not from the residue of chalcocite leaching. Further research and
tronite (Na2Cu(CO3)2•3(H2O)) were also detected. In the precipitate at investigation will be needed to clearly understand the formation me-
35 and 45 °C, the majority of the poorly crystalline material was copper chanism of this secondary covellite. In the glycine-peroxide system, this
iron sulfate. A high sodium phase that is likely natrite (Na2CO3) was secondary covellite was not dissolved but precipitated out, while some
visible in several of the spectra, mixed in with the copper iron sulfate. of the precipitates were dissolved in DI water after harvesting.
Sodium carbonate minerals might be originated from glycine oxidation The precipitation patterns at 35 and 45 °C leaching were totally

76
D. Shin et al. Hydrometallurgy 183 (2019) 71–78

different with the room temperature leaching. It is believed that oxygen


would not be limited because the system was aerated during the test in
higher temperatures, and the oxygen solubility at atmospheric pressure
is 8.7, 6.9, and 6.4 mg/L at 22, 35, and 45 °C, respectively
(U.S._Geological_Survey, 1998). Interestingly, with increasing tem-
perature, iron leachability was increased as shown in Fig. 5 (b). A hy-
pothesis is proposed as follows. By increasing temperature or con-
centration, the solubility of iron would be increased and the dissolved
iron may form copper iron sulfate with copper and sulfate in the so-
lution followed by the precipitation of iron. Iron leaching by the gly-
cine-peroxide system was also tested in this study by using pyrite, and
almost no iron was leached from pyrite. Unfortunately, increase of iron-
glycine complex solubility with temperature has not been reported. The
standard state equilibrium constants (log β) of ferrous-glycinate and
cupric-glycinate were not changed much in the range of 0–50 °C (Shock
and Koretsky, 1995). For copper‑iron-sulfate reaction, very few studies
have been reported only in acidic condition (Gylienė et al., 2010). Cu
(II) complex with tartrate, glycine, and quadrol in the electroless copper
plating solution was reduced by adding reducing agents such as me-
tallic iron. The composition of the precipitate is organic matter, co-
precipitated copper and iron, and the rest of them are some iron hy-
droxyl-compounds and basic iron sulfate. The three different ligands
show different interactions with iron surface, dissolved ferrous and
ferric ions and possible different behavior in co-precipitation. Cu (II)
and Fe (III) form insoluble complex and co-precipitate with organic
ligands such as citric and tartaric acid at alkaline pH while the pre-
cipitation of Fe (II) was very slow at acid pH (Gylienė et al., 1997). The
completeness of precipitation depends on pH and much less on the
metal ion concentration. Another study has been also reported that
ferric ion forms an unstable violet complex by involving linkages
through the sulfur atoms of cysteine (Perrin, 1958).
As stated above, since the precipitates at 35 and 45 °C was more
solid-like and hard to break-up, the precipitation of copper iron sulfate
may inhibit the access of glycine and hydrogen peroxide toward chal-
copyrite surface. In the conventional acidic ferric sulfate chalcopyrite
leaching, as shown in Eq. (1), elemental sulfur is precipitated to form a
passivation layer, however, the results in this study show that some
other mechanisms may be involved in glycine leaching of copper from
chalcopyrite. It is different with the proposed mechanism in the study of
Eksteen et al. (2017), thus more study is needed to identify the detailed
leaching mechanism in various temperature. Although this remains to
be verified through more rigorous study, it is quite clear that under the
conditions used with the specific chalcopyrite concentrates from the
region, increasing temperature has a negative impact on copper ex-
traction by glycine from chalcopyrite, due to copper precipitation and
its passivation effect. Therefore it would be important to develop the
leaching process for minimizing this copper and iron precipitation and
increasing copper leaching efficiency based on these data. More rig-
orous studies are needed to verify which factors affect the precipitation
such as type of oxidants (peroxide or oxygen), oxygen or radical con-
centration, or temperatures and investigate the detailed leaching me-
chanisms of copper from chalcopyrite by glycine-peroxide.

4. Conclusions

The alkaline glycine-peroxide system was used to leach copper from


chalcopyrite concentrate. The addition of glycine and hydrogen per-
Fig. 9. SEM images of the precipitates collected from the leaching liquor at (a) oxide during leaching considerably increased copper extraction. In
room temperature, (b) 35 °C, and (c) 45 °C. The detailed investigations on the addition, copper sulfide and copper sulfate precipitation was detected
red spots are given in Table 3. (For interpretation of the references to colour in in the room temperature leaching liquor, while copper iron sulfate
this figure legend, the reader is referred to the web version of this article.) precipitation was observed in the 35 and 45 °C leaching liquor. This
precipitation may hinder copper extraction rate acting as a possible
passivation layer. The finding in this study is important to understand
the glycine leaching mechanism of copper from chalcopyrite con-
centrate.

77
D. Shin et al. Hydrometallurgy 183 (2019) 71–78

Table 3
The elemental composition of the spots in the SEM images in Fig. 9. Data is presented directly from analysis and not normalized. The total values do not equal 100%
due to surface area roughness. A polished flat sample is ideal to obtain values closer to 100%.
Fig. 9 (a) RT Elemental (%)

Spot # Na S Fe Cu O Total

1 – 15.2 0.15 52.8 – 68.2


2 – 14.4 0.16 44.2 9.3 68.0
3 – 15.9 0.10 52.2 8.6 76.8
4 – 11.4 0.22 62.2 – 73.8
5 4.9 19.2 0.40 57.5 9.6 91.8
6 3.3 14.1 0.46 47.9 8.8 74.5

Fig. 9 (b) 35 °C Elemental (%)

Spot # Na Al Ca K Si S Fe Cu O Total

1 – – 0.20 0.09 1.2 20.7 15.5 39.3 12.5 89.5


2 – – 0.40 0.24 2.8 14.4 13.0 43.9 16.4 91.0
3 4.0 – 0.20 0.18 11.2 8.2 13.1 27.1 23.7 87.7
4 – – 0.30 0.17 3.2 12.9 15.9 36.1 21.9 90.4
5 6.4 1.9 0.80 0.25 9.1 8.0 18.3 23.2 34.5 102.5
6 – – 0.60 0.39 6.1 13.2 18.2 38.4 28.4 105.2

Fig. 9 (c) 45 °C Elemental (%)

Spot # Na Al Ca K Mg Pb Si S Fe Cu O Total

1 8.5 0.69 4.3 0.80 2.6 – 32.6 – 0.29 0.52 40.8 91.1
2 5.4 – 0.92 0.10 – 0.84 3.6 4.3 27.4 16.1 31.4 90.0
3 – – 0.65 0.10 – – 3.6 7.8 20.0 26.7 26.6 85.4
4 – – 0.28 0.10 – – 1.6 10.3 12.2 43.2 6.8 74.5
5 5.2 – 1.29 0.10 – 0.49 2.4 6.6 20.6 22.8 15.6 75.0
6 – – 7.84 – – – 4.5 11.5 12.7 30.4 30.5 97.5

Acknowledgments Gericke, M., Neale, J.W., van Staden, P.J., 2009. A Mintek perspective of the past 25 years
in minerals bioleaching. J. S. Atr. Inst. Min. Metall. 109, 567–585.
Gylienė, O., Šalkauskas, M., Juškėnas, R., 1997. The use of organic acids as precipitants
D. Shin was supported for her sabbatical at the University of Arizona for metal recovery from galvanic solutions. J. Chem. Technol. Biotechnol. 70 (1),
by the Basic Research Project (GP2017-025) of the Korea Institute of 111–115.
Geoscience and Mineral Resources (KIGAM), funded by the Ministry of Gylienė, O., Vengris, T., Nivinskienė, O., Binkienė, R., 2010. Decontamination of solutions
containing Cu(II) and ligands tartrate, glycine and quadrol using metallic iron. J.
Science and ICT of Korea. Hazard. Mater. 175 (1), 452–459.
Hariharaputhiran, M., Zhang, J., Ramarajan, S., Keleher, J.J., Li, Y., Babu, S.V., 2000.
References Hydroxyl radical formation in H2O2-amino acid mixtures and chemical mechanical
polishing of copper. J. Electrochem. Soc. 147 (10), 3820–3826.
Munoz, P.B., Miller, J.D., Wadsworth, M.E., 1979. Reaction mechanism for the acid ferric
Aksu, S., Doyle, F.M., 2001. Electrochemistry of copper in aqueous glycine solutions. J. sulfate leaching of chalcopyrite. Metall. Trans. B 10 (2), 149–158.
Electrochem. Soc. 148 (1), B51–B57. Oraby, E.A., Eksteen, J.J., 2014. The selective leaching of copper from a gold–copper
Baláž, P., Achimovičová, M., Bastl, Z., Ohtani, T., Sánchez, M., 2000. Influence of me- concentrate in glycine solutions. Hydrometallurgy 150, 14–19.
chanical activation on the alkaline leaching of enargite concentrate. Hydrometallurgy Panda, S., Parhi, P.K., Nayak, B.D., Pradhan, N., Mohapatra, U.B., Sukla, L.B., 2013. Two
54 (2), 205–216. step meso-acidophilic bioleaching of chalcopyrite containing ball mill spillage and
Berger, P., Karpel Vel Leitner, N., Doré, M., Legube, B., 1999. Ozone and hydroxyl radicals removal of the surface passivation layer. Bioresour. Technol. 130, 332–338.
induced oxidation of glycine. Water Res. 33 (2), 433–441. Perrin, D.D., 1958. 632. Complex formation between ferric ion and glycine. J. Chem. Soc.
Cheng, C.Y., Lawson, F., 1991. The kinetics of leaching chalcocite in acidic oxygenated (0), 3120–3124.
sulphate-chloride solutions. Hydrometallurgy 27 (3), 249–268. Shock, E.L., Koretsky, C.M., 1995. Metal-organic complexes in geochemical processes:
Córdoba, E.M., Muñoz, J.A., Blázquez, M.L., González, F., Ballester, A., 2008. Leaching of estimation of standard partial molal thermodynamic properties of aqueous complexes
chalcopyrite with ferric ion. Part I: General aspects. Hydrometallurgy 93 (3–4), between metal cations and monovalent organic acid ligands at high pressures and
81–87. temperatures. Geochim. Cosmochim. Acta 59 (8), 1497–1532.
Curreli, L., Garbarino, C., Ghiani, M., Orrù, G., 2009. Arsenic leaching from a gold bearing Tanda, B.C., Eksteen, J.J., Oraby, E.A., 2017. An investigation into the leaching behaviour
enargite flotation concentrate. Hydrometallurgy 96 (3), 258–263. of copper oxide minerals in aqueous alkaline glycine solutions. Hydrometallurgy 167,
Du, T., Vijayakumar, A., Desai, V., 2004. Effect of hydrogen peroxide on oxidation of 153–162.
copper in CMP slurries containing glycine and Cu ions. Electrochim. Acta 49 (25), U.S._Geological_Survey, 1998. Techniques of Water-resources Investigations of the United
4505–4512. States Geological Survey: Chap. A1. Preparations for Water Sampling. U.S.
Duda, L.L., Bartecki, A., 1982. Dissolution of Cu2S in aqueous edta solutions containing Government Printing Office.
oxygen. Hydrometallurgy 8 (4), 341–354. Yu, R.-l., Zhong, D.-l., Miao, L., Wu, F.-d., Qiu, G.-z., Gu, G.-h., 2011. Relationship and
Eksteen, J. and Oraby, E., 2015. Process for selective recovery of chalcophile group effect of redox potential, jarosites and extracellular polymeric substances in bio-
elements, Australian Provisional Patent Application, #2015900865. leaching chalcopyrite by acidithiobacillus ferrooxidans. Trans. Nonferrous Metals
Eksteen, J.J., Oraby, E.A., Tanda, B.C., 2017. A conceptual process for copper extraction Soc. China 21 (7), 1634–1640.
from chalcopyrite in alkaline glycinate solutions. Miner. Eng. 108, 53–66. Zoecklein, B., Fugelsang, K.C., Gump, B., Nury, F.S., 1995. Wine Analysis and Production.
Feng, S., Yang, H., Xin, Y., Gao, K., Yang, J., Liu, T., Zhang, L., Wang, W., 2013. A novel Springer Science & Business Media, New York.
and highly efficient system for chalcopyrite bioleaching by mixed strains of
Acidithiobacillus. Bioresour. Technol. 129, 456–462.

78

You might also like