You are on page 1of 12

Hydrometallurgy 155 (2015) 141–152

Contents lists available at ScienceDirect

Hydrometallurgy

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

Study on the second stage of chalcocite leaching in column with redox


potential control and its implications
Xiaopeng Niu a,b, Renman Ruan a,⁎, Qiaoyi Tan a, Yan Jia a, Heyun Sun a
a
National Engineering Laboratory for Hydrometallurgical Cleaner Production Technology, Institute of Process Engineering, CAS, Beijing 100190, China
b
School of Chemistry and Chemical Engineering, University of Chinese Academy of Sciences, Beijing 100049, China

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

Article history: High-grade natural chalcocite (Cu2S) was leached in column under controlled redox potential (Eh). To satisfy the
Received 8 December 2014 requirements of the shrinking core model, the conditions in the column were kept as uniform as possible by
Received in revised form 19 April 2015 adopting a short height of the middle chalcocite layer, high acid Fe2(SO4)3 feed and high irrigation rate. The ef-
Accepted 26 April 2015
fects of temperature, Fe3+ concentration and Eh on the second stage of chalcocite leaching were investigated.
Available online 30 April 2015
The second stage is sensitive to temperature whereas insensitive to the variations in Eh and Fe3+ concentration
Keywords:
above a certain level. The second stage was divided into two “sub-stages” based on the inflection point at approx-
Chalcocite imately 70% copper dissolution. The excellent linear relationship plotted by the mixed-kinetics indicated that the
Column leaching second stage is controlled by the rates of both chemical reaction and diffusion, in which the limitation of the
Controlled redox potential chemical reaction rate is dominant. The second “sub-stage” (N 70% dissolution) is responsible for the slow kinetics
Shrinking core model of the second stage. Mineralogical study of the residues confirmed that a S0 product layer covered around a
Sulfur product layer shrinking CuS core during the second stage. In addition, high temperature resulted in a more porous S0 layer
which remarkably influences the ion diffusion rate. A dissolution model based on the shrinking core model
was proposed. Implications of the findings for heap bioleaching were discussed.
© 2015 Elsevier B.V. All rights reserved.

1. Introduction CuS þ 2Fe3þ →Cu2þ þ 2Fe2þ þ S0 : ð2Þ

Chalcocite (Cu2S) is the most common secondary copper sulfide


Similar reactions can be written with O2 or Cu2 + as the oxidant.
mineral. Heap bioleaching of low-grade chalcocite ores is one of the
However, Fe3+ in a sulfate medium is the primary oxidant for commer-
most significant industrial applications of bioleaching (Domic, 2007;
cial heap bioleaching of chalcocite in terms of its high efficiency in
Petersen and Dixon, 2007; Watling, 2006). However, current
extracting electrons from a metal sulfide lattice, easy regeneration by
bioleaching operations exhibit low efficiency because of their unsatis-
iron-oxidizing bacteria, and low cost for separating and recovering
factory leaching rate (Bolorunduro, 1999; Brierley, 2008, 2010). Conse-
copper by solvent extraction–electrowinning.
quently, understanding the dissolution mechanism of chalcocite to
The first stage of chalcocite leaching is very fast relative to the sec-
promote copper leaching rate is necessary for commercial application.
ond stage even at ambient temperature. It is controlled by the diffusion
Oxidative dissolution of chalcocite by Fe3+, O2 or Cu2+, either in a
of oxidant to the mineral surface supported by low activation
sulfate or chloride system, occurs in two distinct stages with the produc-
energies(4–25 kJ/mol) and the first-order dependence on Fe3 + or
tion of secondary covellite (Cu1.0–1.2S) as an intermediate (Bolorunduro,
dissolved oxygen concentration (Cheng and Lawson, 1991a; Petersen
1999; Cheng and Lawson, 1991a; Dutrizac and MacDonald, 1974; Mao
and Dixon, 2003).
and Peters, 1983; Mcdonald and Langer, 1983; Miki et al., 2011; Mulak
The second stage is essentially the oxidation of the intermediate or
and Niemiec, 1969; Petersen and Dixon, 2003; Ruan et al., 2013; Ruiz
secondary covellite, which is also described as the “blue-remaining” co-
et al., 1998, 2007; Senanayake, 2007, 2009; Sullivan, 1930; Thomas et
vellite, to produce S0 and Cu2+. The secondary covellite is more reactive
al., 1967):
than the primary mineral (Cheng and Lawson, 1991a, 1991b; King,
1966; Walsh and Rimstidt, 1986). Miki et al. (2011) attributed this effect
Cu2 S þ 2Fe3þ →Cu2þ þ 2Fe2þ þ CuS ð1Þ
to the significantly increased surface area of the secondary covellite
formed by the dissolution of approximately 50% of copper from chalco-
⁎ Corresponding author at: Institute of Process Engineering, Chinese Academy of
cite. The leaching rate of the second stage is more dependent on tem-
Sciences, Beijing 100190, China. perature compared with the first stage, with reported activation
E-mail address: rmruan@ipe.ac.cn (R. Ruan). energies from 55 kJ/mol to 105 kJ/mol (Dutrizac and MacDonald,

http://dx.doi.org/10.1016/j.hydromet.2015.04.022
0304-386X/© 2015 Elsevier B.V. All rights reserved.
142 X. Niu et al. / Hydrometallurgy 155 (2015) 141–152

1974; Petersen and Dixon, 2003). Such high activation energy values 2. Experimental
may suggest that chemical and/or electrochemical reaction controls
the rate of the second stage (Bolorunduro, 1999; Ruan et al., 2013). Con- 2.1. Mineral samples
sequently, the diffusion barrier, which is thought to be due to the forma-
tion of a sulfur product layer, is usually neglected. However, the High-grade natural chalcocite specimen was manually sorted
activation energy value may not be the only criterion that distinguishes from Zijinshan copper ores. The chalcocite sample was prepared by
the rate-determining step. For instance, Ruiz et al. (1998) found that the wet grinding and wet sieving into two narrow size fractions
second stage of digenite leaching in oxygenated chloride media was (− 54 + 30 μm and − 74 + 54 μm). Most of the experiments were
well described by the shrinking core model controlled by diffusion, carried out using 1 g of chalcocite (− 54 + 30 μm size). Results of the
even though apparent activation energy of 84 kJ/mol was obtained. chemical analysis are given in Table 1. Powder XRD confirmed that chal-
Behavior of sulfur in the dissolution of a metal sulfide is an important cocite was the primary phase (Fig. 1). To minimize surface oxidation,
indicator of the mechanism. Acid-soluble metal sulfides such as ZnS, the particle fractions were stored under N2 atmosphere until required
PbS, CuFeS2, CuS and Cu2S are attacked by Fe3 + and H+, resulting in for experiment.
the formation of S0 via the “polysulfide pathway”. More than 90% of
the sulfide moiety of these acid-soluble metal sulfides transforms into 2.2. Apparatus and leaching experiments
S0 by chemical oxidation with FeCl3; this effect is due to the inert prop-
erty of S0 under moderate temperatures and pressures against any abi- The column leaching apparatus is shown in Fig. 2. The leaching ex-
otic oxidation in acidic environments; however, data on CuS and Cu2S periments were conducted using a glass column (6 cm in diameter
are unavailable (Schippers and Sand, 1999). Similarly, the diffusion bar- and 30 cm in length). The column consists of three PVC shells with
rier by the sulfur product layer during the process of leaching metal sul- 5 cm diameter. A layer of glass beads (~5 cm thick) was placed on the
fides, such as ZnS, CuFeS2, CuS and PbS, has been studied extensively top PVC shell to facilitate the distribution of the acid Fe2(SO4)3 lixiviant
(Aydogan et al., 2006; Bobeck and Su, 1985; Cheng and Lawson, over the middle layer of chalcocite. In the middle of the PVC shell, a layer
1991b; da Silva, 2004a, 2004b; Dutrizac, 1989; Fowler and Crundwell, of mixed chalcocite and quartz (~0.5 cm thick) was placed, followed by
1999; Lu et al., 2000; Souza et al., 2007). With regard to chalcopyrite ox- a quartz layer (1 cm thick). In the bottom of the PVC shell, another layer
idation, metal-deficient sulfides, polysulfides, element sulfur and of glass beads (~5 cm thick) was placed to prevent the chalcocite from
jarosites have been proposed to be responsible for the passivation of washing off. The irrigation rate for the column was 5 mL/min. In this
chalcopyrite dissolution (Dutrizac, 1989, 1990; Klauber, 2008; Yang study, the conditions within the column were kept as uniform as possi-
et al., 2015), which poses more to the different reaction pathways ble to satisfy the requirements of the shrinking core model. The condi-
(Senanayake, 2009). The complicated chalcopyrite oxidation may be tions were achieved through a combination of short height of middle
related to critical leaching conditions, particularly the solution Eh chalcocite layer, high acid Fe2(SO4)3 feed and high irrigation rate. The
(Nicol et al., 2010; Velásquez-Yévenes et al., 2010a, 2010b). Unlike design of the column was inspired by the methods reported by Lizama
chalcopyrite, spontaneous oxidation in the second stage of chalcocite (2004).
leaching occurs if the potential for the oxidant couple is greater than The solution Eh was monitored using a FermProbe Pt electrode with
that for the oxidation of the secondary covellite (Miki et al., 2011). Ag/AgCl reference electrode (3.8 M KCl). The pH was measured using a
Chalcocite oxidation always follows the “polysulfide pathway”. METTLER TOLEDO SG8 pH electrode. The temperature of the solution in
Although the chalcocite leaching process appears to be very simple, the round-bottom flask was controlled using an oil bath equipped with
no clear evidence is available on sulfur behavior by both kinetics analy- a digital thermostat and that of the column was controlled using a water
sis and mineralogy. Therefore, a detailed study including clear kinetics bath.
analysis and mineralogy examination would help in elucidating the be- The chemical reagent Fe2(SO4)3 for the kinetic experiments was pre-
havior of sulfur in chalcocite leaching. pared as follows: different amounts of Fe2(SO4)3 and 10 g of 98 wt.%
In general, kinetics of chalcocite has been extensively studied. How- H2SO4 were dissolved in distilled water. Afterward, the solution was
ever, most of the studies that describe the kinetics have focused on adjusted to 1 L.
stirred leaching processes rather than column leaching which is more Approximately 600 mL of the prepared acid Fe2(SO4)3 solution was
similar to industrial heap bioleaching. In addition, very few studies placed into a round-bottom flask and heated to the required tempera-
have been conducted to investigate chalcocite leaching under ture. The solution Eh was controlled within ± 10 mV of the set value
controlled Eh condition (Bolorunduro, 1999; Ruan et al., 2013). Given
that an oxidant is consumed, the Eh decreases from the beginning of
the experiment. Moreover, the solution Eh in the heap is usually main-
tained at a steady level by iron-oxidizing bacteria. Hence, column ♦
∇ SiO2
leaching experiments under controlled Eh can be employed to simulate B ∇
♦ Cu2S
heap bioleaching. ♦

In this work, the effects of temperature and Eh and Fe3+ concentra- ∇ ♦

tion on chalcocite leaching in columns are investigated under controlled ∇ ♦
Intensity

Eh. Eh is controlled by H2O2 instead of KMnO4 to avoid jarosite produc-


tion (Bolorunduro, 1999; Ruan et al., 2013). Scanning electron microsco- ♦
py (SEM)–energy dispersive spectroscopy (EDS), X-ray diffraction (XRD)
and Raman spectroscopy are utilized to characterize the mineralogical A ♦
changes and reveal the behavior of sulfur in chalcocite dissolution. ♦
♦ ♦
∇ ♦

Table 1
Chemical analysis of chalcocite samples (wt.%).
10 20 30 40 50 60 70 80 90
Size fraction (μm) Cu S Fe Si
2θ/°
−74 + 54 71.64 20.95 1.85 0.97
−54 + 30 72.61 22.90 1.67 1.03
Fig. 1. XRD patterns of initial chalcocite samples: (A) −54 + 30 μm and (B) −74 + 54 μm.
X. Niu et al. / Hydrometallurgy 155 (2015) 141–152 143

using 15 wt.% H2O2 solution, which was titrated into the round-bottom 2.4. Raman spectroscopy
flask using an ORP controller connected to a peristaltic pump. The
lixiviant was recycled with the use of a peristaltic pump. For Raman spectroscopy (HORIBA JOBIN YVON ARAMIS), a 633 nm
Aliquots (2 mL) of the reaction solution were sampled periodically line of He–Ne ion laser was used, which was standardized with silicon
using a calibrated pipette. The total Cu and Fe concentrations were mea- before the test. Low intensity laser with 0.04 mW energy was used
sured using inductively coupled plasma optical emission spectrometry throughout the experiment because the samples are easily damaged
(ICP-OES), and the Fe2+ concentration was titrated with potassium di- by energy above 0.1 mW. The Raman data were recorded between
chromate. The concentration of Fe3+ was calculated as the difference 100 and 500 cm−1. Measurements were taken at four to five spots for
between the total and ferrous ion concentrations. The leached residues each sample.
of chalcocite and quartz in the middle of the PVC shell were separated
by wet sieving. The separated chalcocite residue was washed five
times with H2SO4 at pH 1.05 to prevent possible contamination of the 2.5. SEM–EDS
surface by solution species, and then dried under nitrogen atmosphere
for further analysis. Samples were coated with platinum by electro-deposition and ex-
amined by SEM (JSM-7001, Japan) coupled with EDS (INCAX-MAX) at
2.3. XRD 15 kV accelerating voltage.
A portion of the residues was placed into the bottom of the sample
The powder XRD patterns of the residues were collected using a cup, added with epoxy under vacuum, and allowed to harden. The hard-
Rigaku Smartlab diffractometer fitted with a Cu tube and operated at ened epoxy mount was polished, coated with platinum, and then
45 kV and 200 mA. XRD data were collected from 5° to 90° (2θ) in measured by MLA 250 equipped with SEM (Quanta 250) and EDS
0.02° steps at 9°/min speed. (GENESIS).

5
18

xx˚C

xx˚C 1 7
16
9 2
3
4
10

11
15
6
12
xxmL/min

17
14
13
xx mV xx˚C
xx˚C xx˚C

(a) (b)

(c)
Fig. 2. (a) Apparatus employed for column leaching under controlled Eh. (b) The configuration of the column. (c) Bottom of a PVC shell. 1—glass beads, 2—mixed chalcocite and quartz,
3—quartz, 4—glass beads, 5—dripper, 6—pedestal, 7—PVC shell, 8—pore, 9—jacket, 10—ORP probe, 11—round-bottom flask, 12—stirrer, 13—oil bath, 14—ORP digital controller, 15,
16—peristaltic pump, 17—H2O2, 18—water bath.
144 X. Niu et al. / Hydrometallurgy 155 (2015) 141–152

0.9
3. Results and discussion
0.8
3.1. Effect of Fe3+ concentration
0.7
The main purpose of investigating the effect of Fe3+ concentration

Cu Dissolution
0.6
was to check if Fe3 + ion depletion during the reaction could be
neglected. Fig. 3 shows the effect of Fe3+ concentration on copper disso- 0.5 60˚C
lution, with an initial ferric concentration of 1, 3, and 10 g/L respectively. 45˚C
The results show that copper dissolution rate was almost independent 0.4
30˚C
of Fe3 + concentration above 1 g/L. In addition, Fe2 + concentration 25˚C
0.3
was approximately 0.4 g/L in 10 g/L Fe3+ whereas almost undetected
in 1 g/L and 3 g/L Fe3+. This finding confirmed that the criteria of the 0.2
shrinking core model, i.e., constant concentrations of liquid reactants
in solution, were satisfied. 0.1

0.0
0 25 50 75 100 125 150 175 200 225 250
3.2. Effect of temperature
Time(h)
Fig. 4 summarizes the dissolution curves obtained at different tem-
Fig. 4. Effect of temperature on chalcocite dissolution. Conditions: size fraction,
peratures. In general, 45 °C is the distinction of the favorable tempera-
−54 + 30 μm; Fe3+ concentration, 10 g/L; pH, 1.05 ± 0.05; Eh, 750 mV.
ture between the mesophiles and moderate thermophiles, whereas
25 °C is the ambient temperature at which the heap works.
At all temperatures shown in Fig. 4, the leaching reaction had an ev-
ident stage of rapid dissolution, followed by a stage that proceeded at a
slow rate. This observation confirms again that the leaching process than the rest potential of chalcocite reported at approximately 600 mV
proceeds by a two-stage mechanism. Copper was much more prone to (Miki et al., 2011; Petersen and Dixon, 2003).
leaching at high temperatures than it is at low temperatures, especially Figs. 5 and 6 present the effect of Eh on copper dissolution. The
above approximately 45% copper dissolution. This finding shows that slopes of the log rate between 50% and 80% dissolution vs. Eh plot
temperature performs a dominant role in the rate of the second stage showed 0.002 and 0.001 at 30 and 45 °C respectively, which indicated
leaching. The resident time between 70% and 80% copper dissolution that the rate of the second stage leaching was insensitive to Eh. This re-
was almost equal to that between 45% and 70% copper dissolution sult was contrary to that of the study by Bolorunduro (1999), who con-
(Fig. 7). Furthermore, the average dissolution rate between 45% and cluded that Eh governed the rate of the second stage leaching. However,
70% copper dissolution is approximately two fold than that between the experiments were conducted at 75 °C only, thus disregarding the
70% and 80% copper dissolution. This phenomenon would not change variation of temperature on the effect of Eh. Miki et al. (2011) investi-
even though an increase in temperature can cause a significant gated this process in a relatively low range of 500–600 mV in dilute
elevation in copper dissolution rate. A detailed discussion about the chloride solutions at 35 °C; they suggested that oxidation of primary
slow dissolution in the second stage is presented in Section 3.4. or secondary covellite requires at least 600 mV. The dependence of
the rate of chalcocite leaching on Eh, however, in the industrial range,
has not been determined in these studies.
3.3. Effect of Eh
The results of our work confirmed that the rate of the second stage of
chalcocite leaching was insensitive to Eh at moderate temperatures in
Leaching experiments were conducted at various potentials of 650,
the industrial range of 650–800 mV.
700, 750, and 800 mV. The range of Eh was chosen based on heap
bioleaching practices (Ruan et al., 2013), and the values set were greater

0.9
1.0
0.8
0.9

0.8 0.7
Cu Dissolution

0.7 0.6 750mV


Cu Dissolution

0.6 10g/L 700mV


0.5
3g/L 650mV
0.5
1g/L 0.4
0.4
0.3
0.3
0.2
0.2
0.1
0.1
0.0
0.0 0 50 100 150 200 250
0 50 100 150 200 250
Time(h)
Fig. 3. Effect of Fe3+ concentration on chalcocite dissolution. Conditions: size fraction,
−54 + 30 μm; temperature, 30 °C; pH, 1.05 ± 0.05; Eh, 750 mV (throughout this paper Fig. 5. Effect of Eh on chalcocite dissolution. Conditions: size fraction: −54 + 30 μm;
potentials are quoted with reference to the standard hydrogen electrode, SHE). temperature, 30 °C; Fe3+ concentration, 10 g/L; pH, 1.05 ± 0.05.
X. Niu et al. / Hydrometallurgy 155 (2015) 141–152 145

0.9
(Cu1.0–1.2S). The second inflection point at approximately 70% copper
0.8 dissolution corresponds to the noticeable decline in dissolution rate,
which suggests that a change in reaction mechanism may have oc-
0.7 curred. Therefore, in this work, the second stage of chalcocite leaching
was divided into two “sub-stages”: the first “sub-stage” corresponds
Cu Dissolution

0.6 800mV
to 45% to 70% copper dissolution, and the second “sub-stage” corre-
750mV
sponds to N 70% copper dissolution. A part of this work focused on
0.5 700mV
analyzing the kinetics of these two “sub-stages”.
650mV
0.4 The fraction values of the reacted blue-remaining covellite
(Cu1.0–1.2S), X, were tested for the degree of fit to the shrinking core ki-
0.3 netic models given by Eqs. (3) and (4) (Bobeck and Su, 1985; Fowler
and Crundwell, 1999; Levenspiel, 1999; Liddell, 2005; Pritzker, 1996;
0.2
Safari et al., 2009; Su, 1976; Wen, 1968):
0.1
1−ð1−XÞ1=3 ¼ kC t ð3Þ
0.0
0 10 20 30 40 50
1−2=3X−ð1−XÞ2=3 ¼ kP t ð4Þ
Time(h)
MbKCC Cfc 2MbDe Cfs
Fig. 6. Effect of Eh on chalcocite dissolution. Conditions: size fraction, − 54 + 30 μm; Where Kc ¼ KP ¼
temperature, 45 °C; Fe3+ concentration, 10 g/L; pH, 1.05 ± 0.05. ρr0 ρr20

t ¼ t −tC :

3.4. Kinetics analysis As shown in Fig. 8(A–C), neither the chemical reaction control nor
the diffusion control model gave a good linear relationship throughout
Chalcocite or digenite leaching by ferric is widely known to occur in the second stage. In Fig. 8(A), for instance, the data deviated from the
two dramatically different stages. However, two inflection points were excellent linear relationship expected from Eq. (3) after 120 h, which
noticed in the dissolution curves as shown in Fig. 7. The first inflection was close to the time that corresponds to 70% dissolution; subsequently,
point at approximately 45% copper dissolution is consistent with the a distinct decreased rate was observed. However, the excellent linear re-
conversion of all the chalcocite (Cu2S) into blue-remaining covellite lationship obtained from Eq. (4) after 120 h indicated that diffusion

0.9 0.9

0.8
(A) 0.8
(B)
0.7 0.7
Cu Dissolution
Cu Dissolution

0.6 0.6

0.5 0.5

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0.0 0.0
0 25 50 75 100 125 150 175 200 225 250
0 10 20 30 40 50
Time(h) Time(h)

0.9
(C)
0.8

0.7
Cu Dissolution

0.6

0.5

0.4

0.3

0.2

0.1

0.0
0 2 4 6 8 10 12 14
Time(h)

Fig. 7. Dividing the dissolution curves into two “sub-stages” above 45% Cu dissolution. Conditions: size fraction, −54 + 30 μm; Fe3+ concentration, 10 g/L; pH, 1.05 ± 0.05; Eh, 750 mV.
(A) 30 °C; (B) 45 °C; (C) 60 °C.
146 X. Niu et al. / Hydrometallurgy 155 (2015) 141–152

0.30 0.6

(A) (A')
0.25 1/3 0.5

1-(1-X) +B[1-2/3X-(1-X) ]
1-(1-X)

2/3
2/3
2/3
1-2/3X-(1-X)
or 1-2/3X-(1-X) 0.20 2
R =0.995 0.4
1/3
1-(1-X)

KC=0.00164 2
R =0.997
0.15 0.3 k=0.0027
2

1/3
0.10
R =0.996 0.2
KP=3.77193e-4

0.05 0.1

0.00 0.0
0 50 100 150 200 250 0 50 100 150 200 250
Time (h) Time (h)

0.30

(B) 0.5 (B')


0.25 1/3
1-(1-X)

1-(1-X) +B[1-2/3X-(1-X) ]
2/3
2/3
2/3

1-2/3X-(1-X) 0.4
or 1-2/3X-(1-X)

0.20
1/3
1-(1-X)

2
R =0.997 0.3 2
0.15
kC=0.00774 R =0.996
k=0.01259
2

1/3
0.10 R =0.988 0.2

kP=0.00201
0.05 0.1

0.00 0.0
0 10 20 30 40 0 10 20 30 40
Time (h) Time (h)

0.6
0.30
(C) (C')
1-(1-X) +B[1-2/3X-(1-X) ]

1/3 0.5
2/3

0.25 1-(1-X)
2/3
1-2/3X-(1-X)
2/3
or 1-2/3X-(1-X)

0.4
0.20
1/3

2 2
R =0.998
1-(1-X)

R =0.998
0.15 KC=0.0286 0.3 k=0.04809
1/3

2
0.10
R =0.994 0.2
KP=0.00778

0.05 0.1

0.00 0.0
0 2 4 6 8 10 12 0 2 4 6 8 10 12
Time (h) Time (h)

Fig. 8. Plot of data according to the shrinking core model. Conditions: size fraction, −54 + 30 μm; Fe3+ concentration, 10 g/L; pH, 1.05 ± 0.05; Eh, 750 mV. (A, A′) 30 °C; (B, B′) 45 °C; (C, C′)
60 °C.

control became prominent. Similar results observed in Fig. 8(B) and stage, indicating that both chemical reaction and diffusion control con-
(C) further verified the rate-determining step. In general, considering tributed to the dissolution rate throughout the second stage.
just a single control factor throughout the second stage may not be The mixed-model rate constants (Kr) obtained by the application of
reasonable because the relative importance of diffusion and reaction Eq. (5) were used in the Arrhenius plot in Fig. 9, from which, an activa-
control probably varied as the reaction progressed. tion energy of 80.77 kJ/mol can be calculated. This value was found to be
A mixed kinetic equation, i.e., Eq. (5) (Bobeck and Su, 1985; Cordell, within the range of 62.7–104.5 kJ/mol in the ferric sulfate medium
1968; da Silva, 2004b; Liu et al., 2010; Saxena and Mandre, 1992; Su, (Dutrizac and MacDonald, 1974).
1976) was then used to analyze the results: Fig. 10 illustrates the relative contribution of chemical reaction and
  diffusion control on the second stage. The following expressions are
2 plotted against the fractional conversion of blue-remaining covellite, X
Kr t ¼ 1−ð1−XÞ1=3 þ B 1− X−ð1−XÞ2=3 ð5Þ
3 (Bobeck and Su, 1985; Su, 1976):

MbKCC Cfs KCC r0 KC


Where Kr ¼ B¼ ¼ : ð6Þ
ρr0 2De KP
1−ð1−XÞ1=3
The data plotted according to Eq. (5) are shown in Fig. 8(A′–C′). This C¼ h i ð7Þ
1=3
1−ð1−XÞ þ B 1−2=3X−ð1−XÞ2=3
mixed kinetics equation was found to fit the data best for the second
X. Niu et al. / Hydrometallurgy 155 (2015) 141–152 147

-3.0
chemical reaction is dominant. Diffusion control becomes more promi-
nent especially in the second “sub-stage”.
2
-3.5
R =0.9996
k =9.7151
Ea=80.77KJ/mol 3.5. Mineralogy
-4.0
LnK

The changes in the solid phases after dissolution are illustrated by


-4.5 the XRD patterns in Fig. 11. At 60% copper dissolution, the residues
were composed mostly of CuS and S8; Cu2S was undetected. This result
-5.0 suggests the transformation of all Cu2S into CuS after the first stage. In
the wake of the leaching front, as observed in Fig. 11(B–D), the residues
-5.5 were still composed mostly of CuS and S8; however, the intensity of the
diagnostic peak of CuS became weak, whereas that of S8 became strong.
-6.0 Upon further leaching to 90% copper dissolution (Fig. 11(E)), the reac-
tion product was nearly composed completely of S8. The XRD patterns
3.00 3.05 3.10 3.15 3.20 3.25 3.30 provide convincing evidence that the second stage of chalcocite
leaching is essentially the oxidation of the intermediate or secondary
Fig. 9. Arrhenius plot for the second stage of chalcocite leaching. covellite to produce S0.
As shown in Fig. 12, a characteristic of dissolution at different tem-
peratures was that the surfaces of the particles leached at high temper-
atures were much more porous than those leached at low temperatures.
h i The EDS analysis results given in Fig. 12(A′–D′) show that sulfur is
B 1−2=3X−ð1−XÞ2=3
D¼ h i: ð8Þ present on the surfaces as a reaction product. This finding indicates
1−ð1−XÞ1=3 þ B 1−2=3X−ð1−XÞ2=3 that temperature significantly affects the sulfur morphology. The forma-
tion of porous sulfur results in less diffusion barrier than that of thick,
compact sulfur. This finding is consistent with Bolorunduro's (1999)
“C” and “D” represent the percentage of chemical reaction and diffu- assumption that sulfur formed at high temperatures is porous and
sion control, respectively. At 30 °C, the percentage of chemical reaction does not significantly inhibit the second stage.
control is almost equal to that of diffusion control at approximately 80% Raman spectroscopy was used to characterize the surfaces of
dissolution. At 45 and 60 °C, the percentages of diffusion control also leaching residues. In general, Raman spectroscopy provides more sensi-
reached 45% at 80% dissolution. In summary, chemical reaction control tive information on secondary solid phases than XRD (Sasaki et al.,
is dominant during the second stage, but diffusion control becomes 2009) and much more superficial composition information at depth of
progressively prominent especially in the second “sub-stage”. ten or more nanometers compared with EDS. Elemental sulfur was
During the first “sub-stage”, the sulfur product layer may not be suffi- used as standard for Raman spectroscopy. In the Raman spectra
ciently thick to impede the dissolution rate, whereas during the second (Fig. 13), the peaks at 152 and 219 cm−1 are characteristic of elemental
“sub-stage”, the diffusion control becomes prominent because of the in- sulfur caused by the S–S–S bending mode (Holmes and Crundwell,
creasing thickness of the sulfur product layer. This presumption explains 2013; Mycroft et al., 1990; Zhu et al., 1993). The peak at 473 cm−1
the deviation from chemical reaction control after 70% dissolution. could be assigned to either S0 or CuS for their overlapping peaks in the
The above kinetic analysis clearly shows that the second stage of 471–476 cm−1 region (Mernagh and Trudu, 1993; Sasaki et al., 2009;
chalcocite leaching in ferric sulfate medium is controlled by the rates Yang et al., 2015). It is assigned to S0 because it is in proportion to the
of both chemical reaction and diffusion, in which the limitation of the other peaks for S0 and almost coincides with the standard spectra


∇♠ ∇ SiO2 ♠ S8
E ♠♠
1.0 0.0 ♠ ♠ ♠ ♠ ♠♠ ♠ o ♠ ♦ Cu2S o CuS
60˚C ♠ ∇
45˚C D ♠♠
0.9 0.1 ♠ ∇ o♠ ♠ ♠ ♠
o♠

30˚C

Intensity

C ♠
0.8 0.2 ∇ ♠ ♠o ♠ ♠ ♠ ♠ ♠
C D ♠
o ♠
o
B oo
0.7 0.3 ∇ ♠ ∇♠
o o
♠ ♠ ♠♠ ♠ ♦ o


A ♦ ♦
0.6 0.4 ∇ ♦ ♦
∇ ♦

0.5 0.5
0.45 0.50 0.55 0.60 0.65 0.70 0.75 0.80
10 20 30 40 50 60 70 80 90
2θ/°
Cu Dissolution
Fig. 11. XRD patterns of the solid residues. Conditions: 3 g of chalcocite; size fraction,
Fig. 10. Relative contribution of chemical reaction and diffusion control of the second stage − 74 + 54 μm; temperature, 45 °C; Fe3+ concentration, 10 g/L; pH, 1.05 ± 0.05;
of chalcocite leaching. Conditions: size fraction, −54 + 30 μm; Fe3+ concentration, 10 g/L; Eh, 750 mV. (A) Unleached chalcocite, (B) after 60% Cu dissolution, (C) after 70% Cu
pH, 1.05 ± 0.05; Eh, 750 mV. dissolution, (D) after 80% Cu dissolution, and (E) after 90% Cu dissolution.
148 X. Niu et al. / Hydrometallurgy 155 (2015) 141–152

Fig. 12. SEM images of chalcocite particles after leaching at different temperatures. Conditions: size fraction, −54 + 30 μm; Fe3+ concentration, 10 g/L; pH, 1.05 ± 0.05; Eh = 750 mV.
(A) 60 °C after 82% Cu dissolution, (B) 45 °C after 80% Cu dissolution, (C) 30 °C after 81% Cu dissolution, and (D) unleached chalcocite. (A′–D′) EDS spectra of the particles surfaces corre-
sponding to A–D, respectively.

of S0. Furthermore, the weak peak at around 438 cm− 1 is consistent Fig. 14 shows the Backscattered electron (BSE) images of polished
with the weak diagnostic peak of S0 at 438 cm− 1. No peaks for sections of chalcocite particles found in the leached residues. A product
polysulfides were observed, which are expected to be approximately layer was found to cover around a shrinking unreacted core. Spot
at the 452–460 cm−1 region (Mycroft et al., 1990). The Raman spectra analysis was performed by EDS at indicated numbers, as shown in
demonstrate that elemental sulfur is the predominant surface compo- Fig. 14, and the results are provided in Tables 2–5.
nent, which is in agreement with the polysulfide pathway for acid- Element O was not detected in the EDS analysis, which indicates
soluble metal sulfides (Schippers and Sand, 1999). negligible oxidation of the product layer. A S0 product layer clearly
X. Niu et al. / Hydrometallurgy 155 (2015) 141–152 149

(A) increased thickness of the S0 product layer becomes progressively


60˚C @82% prominent.
2. The S0 product layer is almost inert with negligible abiotic oxidation.
45˚C @80% 3. Elevated temperatures significantly accelerate the rate. At Eh
Intensity

N650 mV and Fe3 + concentration N1 g/L, the rate is insensitive to


the variations in Eh and Fe3+ concentration.
30˚C @81% 4. Temperature significantly affects sulfur morphology, which influ-
ences the ion diffusion rate through the S0 product layer.
151 219 5. The diffusion barrier may cause approximately 10%–20% Cu (at.%) to
473
0 remain in the S0 product layer.
S 438

4. Implications for the heap bioleaching of secondary copper sulfide


100 150 200 250 300 350 400 450 500

Raman Shift(cm )
− 1 The results of this research provided some useful pointers to opti-
mize heap bioleaching in practice.
1. The slow rate of the second stage leaching is attributed to the second
(B) “sub-stage” (N70% dissolution) in which the diffusion barrier
45˚C @80%
becomes progressively prominent.
2. Appropriate conditions for the chemical kinetics of copper minerals
are crucial to increase efficiency of heap bioleaching. Elevated tem-
45˚C @70%
Intensity

perature accelerates the chemical reaction rate and causes the forma-
tion of porous S0 product layer, which decreases the prominent
diffusion barrier, especially in the second “sub-stage”.
45˚C @60%
3. The diffusion barrier becomes more prominent because of “particle
151 219 effect”. Consequently, elevated temperature is inadequate for an effi-
473
0 438 cient heap bioleaching because of its limited effect on the accelera-
S
tion of diffusion rate. In addition, above a certain level, the leaching
rate of the second stage is insensitive to the variations in Eh and
100 150 200 250 300 350 400 450 500 Fe3 + concentration. It indicates that above this level, increases in
− 1 Eh and Fe3+ concentration in the heap could no longer provide sig-
Raman Shift(cm ) nificant benefits in terms of copper dissolution rates.
4. Oxidation of S0 product layer by sulfur-oxidizing bacteria results not
Fig. 13. Raman spectra of chalcocite leached (A) at different temperatures with approxi-
only in decreased diffusion barrier but also in the release of remain-
mately 80% Cu dissolution. Conditions: size fraction, −54 + 30 μm; Fe3+ concentration,
10 g/L; pH, 1.05 ± 0.05; Eh, 750 mV; (B) at 45 °C with different % Cu dissolutions. Condi- ing Cu in the layer. The increased thickness of the S0 product layer
tions: 3 g of chalcocite; size fraction, − 74 + 54 μm; Fe3+ concentration, 10 g/L; pH, because of “particle effect” may result in the decline in Fe3+ concen-
1.05 ± 0.05; Eh, 750 mV. tration gradient as well as in diffusive fluxes. The attachment of iron-
oxidizing bacteria on mineral surface enables the direct oxidation of
Fe2+ to Fe3+, thereby avoiding aggregation of Fe2+ on the mineral
surface. The regenerated Fe3+ diffuses through the S0 layer to react
covered around a shrinking CuS core until it disappeared. This finding is at the sulfide–sulfur interface. Furthermore, given the continuous
consistent with the XRD, SEM–EDS and Raman study, which strongly direct oxidation of Fe2 + to Fe3 +, the Fe2 + concentration gradient
supports the kinetics analysis. between the sulfide–sulfur interface and the mineral surface may re-
Cheng and Lawson (1991a) also noticed the decline in leaching rate main at a relatively steady level. The steady concentration gradient is
after 70% dissolution in mixed chloride–sulfate solutions and attributed the driving force for the diffusion of Fe2+ through the S0 layer into
this to the diffusion barrier. However, the authors did not provide suffi- the mineral surface, thereby avoiding aggregation of Fe2 + at the
cient evidence to confirm the presence of S0 product layer. sulfide–sulfur interface.
Moreover, approximately 10%–20% Cu (at.%) remained in the S0
product layer even after repeated washing of the residues. The S0 prod- 5. Conclusions
uct layer not only retards the diffusion of Fe3+ toward the interior core
but also hinders the diffusion of produced Cu2 + toward the surface. The results of this study on column leaching of chalcocite in ferric
Unfortunately, only a few studies have realized the significance of this sulfate medium under controlled Eh led the following conclusions:
phenomenon to heap bioleaching (Dutrizac, 1989). The remaining
1. The slow rate of the second stage leaching is controlled by the rates of
Cu2 + in the S0 product layer coupled with argillic adsorption might
both chemical reaction and diffusion, in which the limitation of the
explain the unheard 90% copper recovery in commercial bioleaching
chemical reaction rate is dominant.
operations as reported by Domic (2007).
2. The second stage is comprised of two “sub-stages”: the first “sub-
stage” (45%–70% dissolution) is predominantly controlled by chemi-
3.6. Proposed dissolution models
cal reactions that occur at the sulfide–sulfur interface; the second
“sub-stage” (N70% dissolution) is responsible for the slow rate of
The dissolution model of the second stage of chalcocite leaching
the second stage leaching because of the prominent diffusion barrier
based on the shrinking core model (Fig. 15) is summarized as
caused by the S0 product layer.
follows:
3. Elevated temperature significantly accelerates the rate of the second
1. The second stage is comprised of two “sub-stages”: the first “sub- stage leaching. At Eh N 650 mV and Fe3+ concentration N1 g/L, the
stage” (45%–70% dissolution) is predominantly controlled by chemi- rate is insensitive to the variations in Eh and Fe3+ concentration.
cal reactions that occur at the sulfide–sulfur interface; in the second 4. A mineralogical study of the residue confirmed that a S0 product
“sub-stage” (N 70% dissolution), the diffusion barrier caused by the layer covered around a shrinking CuS core. The S0 product layer is
150 X. Niu et al. / Hydrometallurgy 155 (2015) 141–152

Fig. 14. BSE images of polished sections of leached chalcocite. Conditions: 3 g of chalcocite; size fraction, −74 + 54 μm; temperature, 45 °C; Fe3+ concentration, 10 g/L; pH, 1.05 ± 0.05; Eh,
750 mV. (A) unleached chalcocite, (B) after 70% Cu dissolution, (C) after 80% Cu dissolution, and (D) after 90% Cu dissolution.

almost inert with negligible abiotic oxidation. Temperature signifi- Acknowledgments


cantly affects the sulfur morphology, which remarkably influences
the ion diffusion rate. The authors acknowledge the funding from IPE, CAS “The introduc-
5. To understand the role of bacteria, an investigation on the bio- tion of outstanding technical talents” program. We thank the reviewers
oxidation of sulfur that covered the mineral surface is being for their enlightening comments and advice.
conducted.

Table 4
EDS analysis of the residue after 80% Cu dissolution (at.%).

Table 2 Number Cu S Cu/S


EDS analysis of unleached chalcocite (at.%). 1 55.65 44.35 1.25
2 50.43 49.57 1.02
Number Cu S Fe Cu/S
3 46.59 53.41 0.87
1 66.21 31.81 1.98 2.08 4 8.73 91.27 0.10
2 65.39 32.06 2.55 2.04 5 12.13 87.87 0.14
3 68.17 31.05 0.78 2.20 6 11.76 88.24 0.13
4 68.52 30.96 0.52 2.21 7 8.50 91.50 0.09
5 68.37 31.02 0.61 2.20 8 19.07 80.93 0.24

Table 5
EDS analysis of the residue after 90% Cu dissolution (at.%).
Table 3
EDS analysis of the residue after 70% Cu dissolution (at.%). Number Cu S Cu/S

1 5.00 95.00 0.05


Number Cu S Cu/S
2 7.70 92.30 0.08
1 44.03 55.97 0.79 3 10.03 89.97 0.11
2 41.65 58.35 0.71 4 8.84 91.16 0.10
3 34.24 65.76 0.52 5 7.31 92.69 0.08
4 23.21 76.79 0.30 6 6.43 93.57 0.07
5 16.24 83.61 0.20 7 6.90 93.10 0.07
6 26.09 73.91 0.35 8 8.86 91.14 0.10
7 25.80 74.20 0.35 9 9.08 90.92 0.10
X. Niu et al. / Hydrometallurgy 155 (2015) 141–152 151

Cu2+
S0
Cu2+ CuS
Boundary layer
Cu+ S-
+ e- e-
Cu+=Cu2 +e-
2+
Fe
2+
Fe
Fe3+
S-=S0+e-

Fe3+
2+
Fe3++e-=Fe
Fig. 15. Schematic diagram of the second stage of chalcocite leaching in ferric sulfate medium. The covellite is referred to as Cu+S− (Goh et al., 2006).

References Lizama, H.M., 2004. A kinetic description of percolation bioleaching. Miner. Eng. 17,
23–32.
Aydogan, S., Ucar, G., Canbazoglu, M., 2006. Dissolution kinetics of chalcopyrite in acidic Lu, Z.Y., Jeffrey, M.I., Lawson, F., 2000. The effect of chloride ions on the dissolution of
potassium dichromate solution. Hydrometallurgy 81, 45–51. chalcopyrite in acidic solutions. Hydrometallurgy 56, 189–202.
Bobeck, G.E., Su, H., 1985. The kinetics of dissolution of sphalerite in ferric chloride solu- Mao, M.H., Peters, E., 1983. Acid pressure leaching of chalcocite. In: Osseo-Asare, K., Miller,
tion. Metall. Trans. B 16B, 413–424. J.D. (Eds.), Hydrometallurgy: Research, Development and Plant Practice. Met. Soc.
Bolorunduro, S.A., 1999. Kinetics of Leaching of Chalcocite in Acid Ferric Sulfate Media: AIME, Warrendale, PA, pp. 243–260.
Chemical and Bacterial Leaching. (M.A.Sc. thesis). University of British Columbia, Mcdonald, G.W., Langer, S.H., 1983. Cupric chloride leaching of model sulfur compounds
Canada. for simple copper ore concentrates. Metall. Trans. B 14B, 559–570.
Brierley, C.L., 2008. A perspective on developments in biohydrometallurgy. Hydrometal- Mernagh, T.P., Trudu, A.G., 1993. A laser Raman microprobe study of some geologically
lurgy 94, 2–7. important sulphide minerals. Chem. Geol. 103, 113–127.
Brierley, C.L., 2010. Biohydrometallurgical prospects. Hydrometallurgy 104, 324–328. Miki, H., Nicol, M., Velásquez-Yévenes, L., 2011. The kinetics of dissolution of synthetic co-
Cheng, C.Y., Lawson, F., 1991a. The kinetics of leaching chalcocite in acidic oxygenated vellite, chalcocite and digenite in dilute chloride solutions at ambient temperatures.
sulphate–chloride solutions. Hydrometallurgy 27, 249–268. Hydrometallurgy 105, 321–327.
Cheng, C.Y., Lawson, F., 1991b. The kinetics of leaching covellite in acidic oxygenated Mulak, W., Niemiec, J., 1969. Kinetics of Cu2S dissolution in acidic solutions of ferric
sulphate–chloride solutions. Hydrometallurgy 27, 269–284. sulphate. Rocz. Chem. 43, 1387–1394.
Cordell, G.B., 1968. Reaction kinetics of the production of ammonium sulfate from anhy- Mycroft, J.R., Bancroft, G.M., McIntyre, N.S., Lorimer, J.W., Hill, I.R., 1990. Detection of
drite. I&EC Process. Des. Dev. 7, 278–285. sulphur and polysulphides on electrochemically oxidized pyrite surfaces by X-ray
da Silva, G., 2004a. Kinetics and mechanism of the bacterial and ferric sulphate oxidation photoelectron spectroscopy and Raman spectroscopy. J. Electroanal. Chem. 292,
of galena. Hydrometallurgy 75, 99–110. 139–152.
da Silva, G., 2004b. Relative importance of diffusion and reaction control during the bac- Nicol, M., Miki, H., Velásquez-Yévenes, L., 2010. The dissolution of chalcopyrite in chloride
terial and ferric sulphate leaching of zinc sulphide. Hydrometallurgy 73, 313–324. solutions: part 3. Mechanisms. Hydrometallurgy 103, 86–95.
Domic, E.M., 2007. A review of the development and current status of copper bioleaching Petersen, J., Dixon, D.G., 2003. The dynamics of chalcocite heap bioleaching. In: Yong, C.A.,
operations in Chile: 25 years of successful commercial implementation. Biomining. Alfantazi, A.M., Anderson, C.G., Dreisinger, D.B., Harris, B., James, A. (Eds.), Hydromet-
Springer-Verlag, Berlin, Heidelberg, New York, pp. 81–96. allurgy 2003: Fifth International Conference in Honor of Professor Ian M. Ritchie.
Dutrizac, J., 1989. Elemental sulphur formation during the ferric sulphate leaching of TMS, pp. 351–364.
chalcopyrite. Can. Metall. Q. 28, 337–344. Petersen, J., Dixon, D.G., 2007. Principles, mechanisms and dynamics of chalcocite heap
Dutrizac, J., 1990. Elemental sulphur formation during the ferric chloride leaching of bioleaching. In: Donati, E.R., Sand, W. (Eds.), Microbial Processing of Metal Sulfides.
chalcopyrite. Hydrometallurgy 23, 153–176. Springer, Dordrecht, The Netherlands, pp. 193–218.
Dutrizac, J.E., MacDonald, R.J.C., 1974. Ferric ion as a leaching medium. Miner. Sci. Eng. 6, Pritzker, M.D., 1996. Shrinking-core model for systems with facile heterogeneous and ho-
59–100. mogeneous reactions. Chem. Eng. Sci. 51, 3631–3645.
Fowler, T.A., Crundwell, F.K., 1999. Leaching of zinc sulfide by Thiobacillus ferrooxidans: Ruan, R.M., Zou, G., Zhong, S.P., Wu, Z.L., Chan, B., Wang, D.Z., 2013. Why Zijinshan copper
bacterial oxidation of the sulfur product layer increases the rate of zinc sulfide disso- bioheapleaching plant works efficiently at low microbial activity — study on leaching
lution at high concentrations of ferrous ions. Appl. Environ. Microbiol. 65, 5285–5292. kinetics of copper sulfides and its implications. Miner. Eng. 48, 36–43.
Goh, S.W., Buckley, A.N., Lamb, R.N., 2006. Miner. Eng. 19, 204–208. Ruiz, M.C., Honores, S., Padilla, R., 1998. Leaching kinetics of digenite concentrate in oxy-
Holmes, P.R., Crundwell, F.K., 2013. Polysulfides do not cause passivation: results from the genated chloride media at ambient pressure. Metall. Mater. Trans. B 29B, 961–969.
dissolution of pyrite and implications for other sulfide minerals. Hydrometallurgy Ruiz, M.C., Abarzúa, E., Padilla, R., 2007. Oxygen pressure leaching of white metal. Hydro-
139, 101–110. metallurgy 86, 131–139.
King, J.A., 1966. Solid state changes in the leaching of copper sulfides. (Doctoral thesis), Safari, V., Arzpeyma, G., Rashchi, F., Mostoufi, N., 2009. A shrinking particle-shrinking core
Imperial College London (University of London). model for leaching of a zinc ore containing silica. Int. J. Miner. Process. 93, 79–83.
Klauber, C., 2008. A critical review of the surface chemistry of acidic ferric sulphate Sasaki, K., Nakamuta, Y., Hirajima, T., Tuovinen, O.H., 2009. Raman characterization of sec-
dissolution of chalcopyrite with regards to hindered dissolution. Int. J. Miner. Process. ondary minerals formed during chalcopyrite leaching with Acidithiobacillus
86, 1–17. ferrooxidans. Hydrometallurgy 95, 153–158.
Levenspiel, O., 1999. Chemical Reaction Engineering. John Wiley & Sons, New York Saxena, N.N., Mandre, N.R., 1992. Mixed control kinetics of copper dissolution for copper
(664 pp.). ore using ferric chloride. Hydrometallurgy 28, 111–117.
Liddell, K.C., 2005. Shrinking core models in hydrometallurgy: what students are not Schippers, A., Sand, W., 1999. Bacterial leaching of metal sulfides proceeds by two indirect
being told about the pseudo-steady approximation. Hydrometallurgy 79, 62–68. mechanisms via thiosulfate or via polysulfides and sulfur. Appl. Environ. Microbiol.
Liu, X.Y., Wu, B., Chen, B.W., Wen, J.K., Ruan, R.M., Yao, G.C., Wang, D.Z., 2010. Bioleaching 65, 319–321.
of chalcocite started at different pH: response of the microbial community to Senanayake, G., 2007. Chloride assisted leaching of chalcocite by oxygenated sulphuric
environmental stress and leaching kinetics. Hydrometallurgy 103, 1–6. acid via Cu (II)–OH–Cl. Miner. Eng. 20, 1075–1088.
152 X. Niu et al. / Hydrometallurgy 155 (2015) 141–152

Senanayake, G., 2009. A review of chloride assisted copper sulfide leaching by oxygenated Glossary
sulfuric acid and mechanistic considerations. Hydrometallurgy 98, 21–32.
Souza, A.D., Pina, P.S., Leão, V.A., Silva, C.A., Siqueira, P.F., 2007. The leaching kinetics of a b: stoichiometric coefficient
zinc sulphide concentrate in acid ferric sulphate. Hydrometallurgy 89, 72–81. B: constant
Su, H., 1976. The Kinetics of Dissolution of Sphalerite in Ferric Chloride Solution. (PhD Cfc: Fe3+ concentration at the surface of the core (mol m−3)
Thesis). University of Idaho, Moscow. Cfs: Fe3+ concentration at the surface of the particle (mol m−3)
Sullivan, J.D., 1930. Chemistry of Leaching Chalcocite. US Bureau of Mines, Washington, De: effective diffusivity of sulfur layer (m2 s−1)
DC, USA, p. TP 473. t: the time for leaching blue-remaining covellite (s)
Thomas, G., Ingraham, T., MacDonald, R.J.C., 1967. Kinetics of dissolution of synthetic t⁎: the total leaching time (s)
digenite and chalcocite in aqueous acidic ferric sulphate solutions. Can. Metall. Q. 6, tc: the time for the conversion of chalcocite into blue-remaining covellite (s)
281–292. KC: kinetic parameter for reaction control (s−1)
Velásquez-Yévenes, L., Nicol, M., Miki, H., 2010a. The dissolution of chalcopyrite in chlo- KCC: chemical reaction rate constant (m s−1)
ride solutions: part 1. The effect of solution potential. Hydrometallurgy 103, 108–113. KP: kinetic parameter for product diffusion control (s−1)
Velásquez-Yévenes, L., Miki, H., Nicol, M., 2010b. The dissolution of chalcopyrite in chlo- Kr: mixed-model rate (s−1)
ride solutions: part 2: effect of various parameters on the rate. Hydrometallurgy M: molecular weight of particular sulfide mineral (kg mol−1)
103, 80–85. r0: radius of original particles (m)
Walsh, C.A., Rimstidt, J.D., 1986. Rates of reaction of covellite and blaubleibender covellite X: the fraction of reacted blue-remaining covellite
with ferric iron at pH 2.0. Can. Mineral. 24, 35–44. X⁎: the fraction reacted
Watling, H.R., 2006. The bioleaching of sulphide minerals with emphasis on copper sul-
phides — a review. Hydrometallurgy 84, 81–108.
Wen, C.Y., 1968. Noncatalytic heterogeneous solid fluid reaction models. Ind. Eng. Chem. Greek letters
60, 34–56.
Yang, Y., Liu, W.H., Chen, M., 2015. XANES and XRD study of the effect of ferrous and ferric ρ: density of sulfide mineral (kg m−3)
ions on chalcopyrite bioleaching at 30 °C and 48 °C. Miner. Eng. 70, 99–108.
Zhu, X.M., Li, J., Bodily, D.M., Wadsworth, M.E., 1993. Transpassive oxidation of pyrite.
J. Electrochem. Soc. 140, 1927–1935.

You might also like