ArticlePDF Available

A case study of flotation of chalcopyrite-molybdenite ore using a mixture of pyrogallic acid and o-isopropyl ethylthiocarbamate: Insights from ab initio calculations

Authors:

Abstract

To date, lime-free flotation processes are increasingly advocated due to many problems caused by the traditional process using lime (CaO) as a depressant and xanthate (BX) as a collector. In this work, we developed a combination of pyrogallic acid (PGA) and o-isopropyl ethylthiocarbamate (Z200) as an eco-friendly alternative to the lime-xanthate process for chalcopyrite-molybdenite (ccp-mlb) flotation. Roughing flotation experiments showed that this process yields ccp-mlb concentrate with similar recovery and grade of Cu and Mo as in the traditional lime-xanthate process. However, the huge advantage of the novel process is that it requires ten times lower dosages of sodium sulfide in the subsequent separation of ccp-mlb concentrate. The ab initio calculations explained this trend at the atomistic level, where hydrogen sulfide ions could desorb collector more easily from the Z200@ccp complex compared with BX@ccp. The same method could be applied for environmentally friendly separations of other resources with similar composition.
A case study of flotation of chalcopyrite-molybdenite ore using a mixture
of pyrogallic acid and o-isopropyl ethylthiocarbamate: Insights from
ab initio calculations
Yuanjia Luo
a
, Leming Ou
a
, Wei Sun
a
, Haisheng Han
a
, Jianhua Chen
b
, Jian Peng
a,
a
School of Minerals Processing and Bioengineering, Central South University, Changsha 410083, China
b
School of Resources, Environment and Materials, Guangxi University, Nanning 530004, China
article info
Article history:
Received 10 February 2023
Received in revised form 31 March 2023
Accepted 5 April 2023
Keywords:
Flotation
Chalcopyrite-molybdenite
Novel process
Sodium sulfide
Desorb
abstract
To date, lime-free flotation processes are increasingly advocated due to many problems caused by the
traditional process using lime (CaO) as a depressant and xanthate (BX) as a collector. In this work, we
developed a combination of pyrogallic acid (PGA) and o-isopropyl ethylthiocarbamate (Z200) as an
eco-friendly alternative to the lime-xanthate process for chalcopyrite-molybdenite (ccp-mlb) flotation.
Roughing flotation experiments showed that this process yields ccp-mlb concentrate with similar recov-
ery and grade of Cu and Mo as in the traditional lime-xanthate process. However, the huge advantage of
the novel process is that it requires ten times lower dosages of sodium sulfide in the subsequent separa-
tion of ccp-mlb concentrate. The ab initio calculations explained this trend at the atomistic level, where
hydrogen sulfide ions could desorb collector more easily from the Z200@ccp complex compared with
BX@ccp. The same method could be applied for environmentally friendly separations of other resources
with similar composition.
Ó2023 The Society of Powder Technology Japan. Published by Elsevier B.V. and The Society of Powder
Technology Japan All rights reserved.
1. Introduction
Porphyritic copper-molybdenum ores are the main mining
resource for chalcopyrite and pyromorphite, which account for
about 50% of global production of pyromorphite and 75% of global
production of copper [1,2]. Due to the excessive use of lime, which
has led to issues like pipeline jamming, environmental disruption,
poor selectivity, foam viscosity, difficult recovery of precious met-
als, and treatment with pyrite in dressing plants [3–5], the tradi-
tional process using lime (CaO) as a depressant and xanthate
(BX) as a collector is becoming increasingly limited. As a result,
lime-free flotation processes are gaining more and more attention.
Therefore, a new reagent scheme is urgently needed to be found to
separate the chalcopyrite and pyromorphite.
The beneficiation of copper-molybdenum ores is generally car-
ried out by a two-stage flotation method. The first stage uses thiol
collectors to recover copper sulfide and pyromorphite as a mixed
concentrate, and the second stage uses collectors to selectively
recover copper sulfide and pyromorphite [6,7]. Generally, in the
second stage but due to the similar floatability of pyromorphite
and chalcopyrite, their separation efficiency is low [8,9]. Over the
decades, researchers have developed a variety of inorganic and
organic reagents to inhibit copper sulfide in the flotation of
pyromorphite. Examples include lignosulphonates [10], disodium
carboxymethyltrithiocarbonate [11], pseudo-glycolythiourea acid
[12], L-cysteine [13], thiocarbonohydrazide [14], poly
(acrylamide-allylthiourea) (PAM-ATU) [15], O-carboxymethyl chi-
tosan [16], rhodanine-3-acetic acid [17], etc. However, these
agents have been studied almost exclusively in laboratory experi-
ments and have not been widely used on an industrial scale, and
the NaHS and Na
2
S are still the commonly used copper inhibitors
[18].
In this paper, we describe the flotation separation of pyrite-
containing porphyry Cu-Mo sulfide ore from Dexing, Jiangxi pro-
vince. In the first stage, we selected a suitable, environmentally
friendly reagent to depress the floatability of pyrite and provide
maximum recovery of chalcopyrite and Mo-rich minerals. In the
second stage, our focus was on improving the efficiency of sodium
sulfide (Na
2
S) to remove the collector from the chalcopyrite surface
and, eventually, to reduce Na
2
S dosage. Herein, we developed a
new process for the instant, environmentally friendly flotation of
chalcopyrite/molybdenite ore that combines PGA, Z200, and
Na
2
S. Compared with the traditional process that utilizes lime
https://doi.org/10.1016/j.apt.2023.104039
0921-8831/Ó2023 The Society of Powder Technology Japan. Published by Elsevier B.V. and The Society of Powder Technology Japan All rights reserved.
Corresponding author.
E-mail address: peng.jian@csu.edu.cn (J. Peng).
Advanced Powder Technology 34 (2023) 104039
Contents lists available at ScienceDirect
Advanced Powder Technology
journal homepage: www.elsevier.com/locate/apt
(CaO) as the inhibitor and butyl xanthate (BX) as the collector, the
proposed treatment was more efficient and required 70% less
Na
2
S to desorb the collector from copper concentrates. The key
intermolecular interactions responsible for lower consumption of
Na
2
S in instant flotation were revealed by ab initio calculations.
The combination of reagents proposed in this work is characterized
by low consumption of chemicals, low cost, and reduced environ-
mental pollution, and represents a promising separation strategy
that can be applied to similar mineral resources in the future.
2. Materials and methods
2.1. Minerals and reagents
2.1.1. Minerals
The studied pyritic Cu-Mo sulfide ore came from Dexing, Jiangxi
Province, China. The samples were submitted to a sequence of
coarse, medium, and fine crushing (Fig. S1,Supplementary
material), and the particles smaller than 2 mm were collected for
flotation experiments and mineralogical analysis. Samples were
stored in sealed bags at –4°C under a controlled atmosphere to
prevent oxidation.
The Cu, Mo, and S content in samples before and after flotation
was determined by ICP-MS. Other elements were quantified by
XRF. Table 1 presents the chemical analysis of the ore reveals. As
presented in Table 1, the content of the four most abundant ele-
ments is 0.37% Cu, 0.026% Mo, 2.52% S, and 2.79% Fe. The amounts
of other elements are too low to be utilized comprehensively. Min-
eralogical phase analysis was carried out on the ore samples using
a polarizing microscope. The mineral composition of the examined
ore is listed in Table 2. As presented in Table 2, the useful minerals
in this ore are mainly chalcopyrite (CuFeS
2
) and molybdenite
(MoS
2
), and their relative mineral contents are 1.28% and 0.054%,
respectively. Other metal sulfide minerals include pyrite with a rel-
ative content of 2.14% and trace amounts of galena and sphalerite.
Gangue minerals account for about 95.66%, mainly quartz, feldspar,
muscovite, biotite, chlorite, etc. The optical microscope image of
the raw ore is shown in Fig. 1. There are three main inlaid forms
of chalcopyrite. First, chalcopyrite is packed in quartz geodes of
quartz veins, and the grain shape and size are limited by the geodes
(Fig. 1(a)). Second, chalcopyrite is filled and metasomatized along
the edge of the pyrite, showing an irregular granular distribution
(Fig. 1(b) and 1(c)). Third, molybdenite is filled along the edge or
gap of chalcopyrite, and the two were closely associated (Fig. 1
(d) and 1(e)). There are also three main embedded forms of molyb-
denite in the ore. First, molybdenite fills the gaps in the quartz
grains or in the quartz geodes (Fig. 1(f) and 1(g)). Second, molyb-
denite is disseminated in gangue, and molybdenite is sometimes
seen contained in pyrite crystals (Fig. 1(h)). Third, molybdenite fills
along with ore fractures and replaces chalcopyrite (Fig. 1(i)).
2.1.2. Reagents
Butyl xanthate (BX, 92%), o-isopropyl ethylthiocarbamate
(Z200, 93%) and kerosene (95%) used as collectors in flotation
experiments were purchased from Zhuzhou Chemical Plant, China.
Terpineol (95%) was obtained from Guangxi Long Chemical Co.,
Ltd., China and used as a foaming agent. The pH was adjusted by
CaO (p.a., Aladdin Reagent Co., Ltd., Shanghai, China) and Na
2
CO
3
(p.a., Tianjin Kemio Chemical Reagent Co., Ltd.). Pyrogallic acid
(95%) and Na
2
S (95%) used as inhibitors were purchased from
Shanghai Mailin Biochemical Co., Ltd., China. All solutions were
prepared on the day of the experiment. Tap water was used for
all tests.
2.2. Flotation experiments
The flotation experiments on a laboratory scale were done on
XFG-type flotation device (Jilin Exploration Machinery Plant,
China) equipped with a 3.0 L flotation cell. The rotor speed was
1500 rpm. For each flotation test, 1000 g of ore was mixed
with 2.0 L of tap water in a flotation cell and stirred for 3 min
to form a homogeneous slurry with a grind fineness of 0.074 mm
(70% particles). Flotation tests were performed at 25 °C. The con-
centrates were separated by filtration, dried and weighed, and then
analyzed for the content of Cu, Mo and S. The recoveries were cal-
culated from the distribution of elements in concentrate and tail-
ing and the grade of the product. The amount of Cu, Mo, and S
was used to represent the grade and recovery of ccp, mlb and
pyr, respectively. Each flotation experiment was performed in trip-
licates, and final values were reported as means where error bars
represent standard deviations.
2.3. Ab initio calculations
Ab initio quantum chemistry calculations were performed with
the CASTEP module in BIOVIA Materials Studio to compare the
adsorption affinity of hydrogen sulfide ion, Z200 and BX on ccp
(112) surface. The GGA algorithm with PW91 functional was
selected [19–22]. During the calculation, TS method for dispersion
correction were added [21–23], and ultrasoft pseudopotentials
were adopted [24–31]. To model chalcopyrite, the cutoff energies
(E
cut-off
) was set to 351 eV and the quality of k-points was set to
fine [19]. Medium quality was selected for geometry optimization
in CASTEP.
The structure of ccp crystal cell (Fig. 2(a)) was retrieved from
American Mineral Crystallographers Database (AMCSD) [32–34].
The (112) surface of ccp was split from the crystal cell since it
has low surface energy [35] and packed into 2 21 supercell
surrounded by 30 Å of vacuum. The optimized structure of con-
structed ccp (112) surface is shown in Fig. 2(b). The 3D structures
of Z200, HS
, and BX were pre-optimized at k-point of gamma in a
30 30 30 Å cell and then allowed to interact with ccp surface.
Their configurations in the adsorbed state are shown in
Table 1
The elemental composition of ores (mass %).
Element Cu Mo S Fe Pb Zn Cl Au(g/t) Rb Sr
Content 0.37 0.026 2.52 2.79 0.028 0.037 0.027 0.11 0.023 0.052
Element Ag(g/t) Ba Ce Na
2
O MgO Al
2
O
3
SiO
2
K
2
O CaO P
2
O
5
Content 0.04 0.048 0.054 4.49 1.97 15.48 62.61 5.02 3.96 0.2
Table 2
The mineral composition of raw ore sample (mass %).
Minerals Content Minerals Content
Chalcopyrite 1.28 Muscovite 5.43
Molybdenite 0.054 Biotite 5.18
Pyrite 2.14 Calcite 0.18
Sphalerite 0.009 Epidote 0.015
Galena 0.005 Chlorite 0.99
Quartz 27.37 Apatite 0.33
Feldspar 56.16
Y. Luo, L. Ou, W. Sun et al. Advanced Powder Technology 34 (2023) 104039
2
Fig. 3(a)-(c). The two surface ccp layers were optimized while the
other four bulk layers were restricted.
The adsorption energy (
D
E) was obtained by the following
formula:
D
E¼Eðccp surface þadsorbateÞEðccp surfaceÞEðadsorbateÞ
ð1Þ
The more negative
D
Evalues indicate stronger interaction
between the ccp surface and the adsorbate.
As chalcopyrite is a sulfide mineral and is relatively hydropho-
bic, the calculations were carried out under vacuum conditions.
3. Results and discussion
3.1. Roughing flotation of chalcopyrite-molybdenite ore
Flotation efficiency of chalcopyrite (ccp) and molybdenite (mlb)
was improved by grounding the samples to obtain fine-grained
particles (Fig. 4). Fig. 4(a) and 4(b) illustrate the traditional
chalcopyrite-molybdenite-pyrite (ccp-mlb-pyr) bulk flotation and
the new ccp-mlb instant flotation, respectively. Fig. 4(c) shows
the process flow applied in the second step of ccp/mlb flotation
separation test.
In this study, we applied these two approaches to recover chal-
copyrite (ccp) and molybdenite (mlb) from ccp-mlb ore. In a con-
ventional approach, ccp and mlb were recovered simultaneously
using sodium carbonate as pH regulator, CaO as the inhibitor and
butyl xanthate (BX) as the collector. The efficiency of the tradi-
tional approach was compared with the novel, instant flotation
process that utilizes pyrogallic acid (PGA) as the inhibitor and
o-isopropyl ethylthiocarbamate (Z200) as the collector. Both
approaches require the addition of sodium sulfide as the ccp flota-
tion inhibitor to separate ccp from mlb.
The results in Fig. 5(a) and Fig. 5(b) show the influence of CaO
and BX dosage on the grade of Cu and Mo in ccp-mlb bulk concen-
trates and S in pyrite (pyr) concentrate. The corresponding raw
Fig. 1. Optical microscopy and SEM images of the untreated ore.
Fig. 2. The crystal cell of chalcopyrite (a) and the optimized (112) surface of
chalcopyrite (b).
Y. Luo, L. Ou, W. Sun et al. Advanced Powder Technology 34 (2023) 104039
3
data are given in Table S1 and Table S2. An increase in CaO dosage
up to 2000 g/t improved the grade of Cu and Mo reaching 10.88%
and 0.7%, respectively. The grade of Cu was almost independent
of CaO dosage, while the grade of Mo decreased at dosages above
2000 g/t. This result indicates that the optimum CaO dosage for
Mo recovery from ccp-mlb ore is 2000 g/t.
Keeping the amount of lime fixed at 2000 g/t, we studied the
influence of BX dosage on the recovery and grade of ccp-mlb con-
centrate. At 60 g/t BX, the recoveries increased to 82.05% and
75.12%, respectively. Further increase in BX negligibly influenced
the recovery of the minerals from ore. Corresponding grades of
Cu and Mo were 11% and 0.7%, respectively. Moreover, the
recovery and grade of S in pyr concentrate were 67% and 25%.
In a new approach, we have used environmentally friendly PGA
and Z200 as replacements for CaO and BX. The results in Fig. 5(c)
and Fig. 5(d) show the influence of reagent dosage on the recovery
and Cu-Mo grade in ccp-mlb bulk concentrate and S in pyr concen-
trate. The raw data can be found in Table S3 and Table S4. The
recoveries of two transition metals in bulk concentrate increased
with the increase in PGA dosage up to 300 g/t. A higher amount
of PGA resulted in a slight recovery drop. Furthermore, the influ-
ence of Z200 dosage was studied by keeping the amount of PGA
at this optimum level. Fig. 5(d) shows that the recovery of Cu
and Mo increases up to 30 g/t Z200, and remains constant at higher
dosages. Under optimum reagent dosage (300 g/t PGA and 30 g/t
Z200), the recoveries of Cu and Mo in ccp-mlb bulk concentrate
reached 87% and 89% with grades of 12% and 0.8%. More-
over, the recovery and grade of S in pyr concentrate were 68%
and 33%.
Comparing the results of the first step of conventional (A) and
instant ccp-mlb bulk flotation (B), it is clear that the dosages of
inhibitor and collector are significantly lower for the instant flota-
tion process (2000 g/t and 60 g/t of CaO and BX vs. 300 g/t and 30 g/
t of PGA and Z200). The Cu and Mo recoveries in concentrate B
were higher by 5% and 3%. The Cu grade was also slightly
improved in concentrate B while Mo grade was similar in both con-
centrates. The recovery of pyr concentrate was increased by 1% in
an instant flotation, and the S grade was improved by 8%. We con-
clude that the combination of novel, eco-friendly flotation reagent
results in the efficient separation of ccp-mlb and pyr.
The addition of sodium sulfide in a bulk concentrate effectively
depresses chalcopyrite flotation and allows the separation of chal-
copyrite from molybdenite in ccp-mlb bulk concentrate. In this
study, we optimized the dosage of Na
2
S to maximize recovery
and grade of the minerals. The results of ccp-mlb bulk separation
were compared for traditional flotation (concentrate A, Fig. 6(a))
and a new, instant flotation scheme (concentrate B, Fig. 6 (b)).
The raw experimental data can be found in Table S5 and Table S6.
Fig. 3. The optimized geometries of Z200 (a), hydrogen sulfide ion (b), and BX (c).
Fig. 4. The process scheme illustrating (a) traditional ccp-mlb-pyr rough flotation; (b) newly proposed ccp-mlb instant flotation and (c) flotation separation of ccp and mlb.
Y. Luo, L. Ou, W. Sun et al. Advanced Powder Technology 34 (2023) 104039
4
The results in Fig. 6(a) show that the grade and recovery of both
minerals increased up to 3000 g/t Na
2
S, while higher concentra-
tions have negligible impact on the separation efficiency. The effec-
tive separation of concentrate was confirmed by high ccp recovery
(88.49%) with good Cu grade (10.99%), as well as with excellent
mlb recovery and Mo grade (89.12% and 5.74 %, respectively).
The trends in recovery and grade of both minerals from con-
centrate B (Fig. 6(b)) were consistent with Fig. 6(a). However, only
300 g/t Na
2
S was necessary to reach the same separation effi-
ciency as in concentrate A. The corresponding recovery and Cu
grade of ccp were 88.03% and 11.87%, while that of mlb and Mo
grade were 90.85% and 6.08%. These results indicate that the
selection of flotation inhibitor and collector applied in the first
step of ccp-mlb flotation have a significant influence on the
dosage of chalcopyrite depressant required for the efficient sepa-
ration of bulk concentrate in the second step. Instant bulk flota-
tion with eco-friendly PGA and Z200 requires only one-tenth of
Na
2
S dosage from that of the conventional lime-xanthate process
in the subsequent separation step.
3.2. Closed-circuit experiments
The type of the reagents and their dosages optimized in rough-
ing flotation experiments were taken as the starting point for
closed-circuit flotation tests. The experimental procedure for this
test is illustrated in Fig. 7. In the first step, Cu-Mo concentrate A
was obtained by instant flotation using the combination of
Z200 + PGA. In the cleaning stage, only PGA inhibitor was used.
Fig. 5. Grades and recoveries of Cu, Mo in ccp-mlb bulk concentrate and S in pyr concentrate as a function of CaO (a) and BX (b), PGA (c) and Z200 (d) dosages.
Fig. 6. The effect of Na
2
S dosage on the separation efficiency of ccp-mlb bulk concentrate obtained by (a) conventional lime-xanthate process and (b) instant PGA-Z200
flotation.
Y. Luo, L. Ou, W. Sun et al. Advanced Powder Technology 34 (2023) 104039
5
Afterward, asynchronous flotation according to a conventional
lime-xanthate scheme gave Cu-Mo concentrate B. Since the con-
ventional separation of Cu-Mo/S requires large amount of lime rel-
atively large loss of Mo is expected. As shown in 3.1, two kinds of
Cu-Mo mixed concentrates have a similar grade and recovery of Cu
and Mo but concentrate A requires a significantly lower amount of
sodium sulfide to desorb Cu from the collector. Using the flow sys-
tem illustrated in Fig. 7 it is possible to adjust the system to obtain
as many instant flotation Cu-Mo concentrates as possible, which is
beneficial to the recovery of Mo.
Under the closed-circuit experimental setup (Fig. 8) and using
the same scavenger and cleaner as in roughing separation of ccp
and mlb from ccp/mlb concentrate it was possible to halve the
amounts of reagents. Moreover, sodium sulfide dosage for Cu-Mo
concentrate A (Fig. 8(a)) was only one-tenth of the dosage for
Cu-Mo concentrate B (Fig. 8(b)). The final yields of three minerals
obtained by closed-circuit flotation are listed in Table 3. High
recoveries of Cu and Mo from corresponding concentrates
(89.81% and 80.50%) and excellent grades (23.40% and 45.50%,
respectively) have confirmed that the newly developed reagent
scheme efficiently separates ccp from mlb. This is provided by
closed-circuit experiments designed to exploit the fact that the
concentrates obtained by traditional and instantaneous flotation
require different amounts of sodium sulfide for collector desorp-
tion from copper concentrates.
3.3. Ab initio calculations
Ab initio quantum chemical calculations provided insights into
the adsorption of Z200 and BX on chalcopyrite (112) surface and
the affinity of hydrogen sulfide ions (HS
) to desorb collectors from
the surface. Table 4 lists the adsorption energies (
D
E) between
Z200 and BX and ccp surface, as well as
D
Ebetween four HS
and the adsorption complexes of ccp with Z200 and BX. The lower
D
Efor the adsorption of HS
ions on the surface of Z200@ccp
Fig. 7. Experimental procedure for closed-circuit Cu-Mo flotation.
Y. Luo, L. Ou, W. Sun et al. Advanced Powder Technology 34 (2023) 104039
6
complex confirms a higher affinity of sodium sulfide to desorb
Z200 and lower dosages of reagent needed for separation of ccp
and mlb from concentrate.
The optimized configurations for the adsorption of hydrogen
sulfide ions to Z200@ccp and BX@ccp complexes are shown in
Fig. 9 (a) and Fig. 9 (b). The binding of HS
ions to Z200@ccp cre-
ated a Cu-S bond length of 4.64 Å. Accordingly, the lengths of
Cu-S bonds of BX@ccp became 4.76 Å and 3.90 Å. These Cu-S dis-
tances are far larger than the covalent radius of Cu-S, thus Cu-S
bonds all have broken down. This result suggests that both collec-
tors could be efficiently desorbed by hydrogen sulfide ions formed
as a product of Na
2
S hydrolysis.
Fig. 8. Experimental procedure for closed-circuit Cu/Mo separation.
Y. Luo, L. Ou, W. Sun et al. Advanced Powder Technology 34 (2023) 104039
7
Fig. 10 (a) and Fig. 10 (b) show the electron density differences
for Z200@ccp and BX@ccp complexes after the adsorption of four
HS
ions. The regions with blue color are parts of the adsorption
complex with the increased electron density compared with
pristine adsorbent, while red-colored regions are the ones with
the decreased electron density. As presented in Fig. 10 (a) and
Fig. 10 (b), this electron transfers in the direction of Cu-S were
not observed in case of HS
adsorption to Z200@ccp and BX@ccp
complexes, which indicates that both collectors were successfully
desorbed from ccp (112) surface.
In summary, ab initio calculations explained the relative affinity
of HS
ions to desorb Z200 and BX collectors from ccp on an atomic
level. This result provides insights into the molecular mechanisms
of instant flotation and explains why lower quantities of Na
2
S are
required to desorb collector from Z200@ccp complex.
4. Conclusions
In this paper, we applied an eco-friendly combination of pyro-
gallic acid (PGA) and o-isopropyl ethylthiocarbamate (Z200) to
replace the traditionally used lime-xanthate process. To efficiently
separate ccp from mlb, we used sodium sulfide as the inhibitor.
Roughing and closed-circuit experimental setups were applied to
separate pyrite-rich CuS/MoS ore. Compared with ccp-mlb concen-
trate obtained by conventional process, concentrate obtained by
instant PGA-Z200 flotation required significantly lower dosages
Table 3
The results of closed-circuit flotation experiments (given in mass %).
Product Yield Grade Recovery
Cu Mo S Cu Mo S
Chalcopyrite concentrate 1.42 23.40 0.25 23.98 89.81 13.65 13.51
Molybdenite concentrate 0.046 0.78 45.50 32.75 0.097 80.50 0.60
Pyrite concentrate 4.31 0.38 0.0064 46.75 4.43 1.06 79.96
Tailings 94.224 0.022 0.0013 0.16 5.67 4.79 5.93
Raw ore 100.00 0.37 0.026 2.52 100.00 100.00 100.00
Table 4
Interaction energies between Z200 and BX collectors and ccp (11 2) surface, and the
interaction energies between HS
ions and Z200@ccp and BX@ccp complexes.
Mineral surfaces (Adsorbent) Adsorbate
D
E(KJ/mol)
ccp Z200 56.60
BX 157.00
Z200@ccp 4 HS
446.65
BX@ccp 4 HS
310.81
Fig. 9. The optimized configuration of Z200@ccp (a) and BX@ccp (b) complexes upon binding of four hydrogen sulfide ions.
Fig. 10. The difference in electron densities after four HS
ions on Z200@ccp complex (a) and four HS
ions on BX@ccp complex (b).
Y. Luo, L. Ou, W. Sun et al. Advanced Powder Technology 34 (2023) 104039
8
of Na
2
S to reach the same separation efficiency. The closed-circuit
setup was designed to exploit this difference, which resulted in
excellent recoveries and grades of ccp and mlb (89.81% with
23.40% Cu, and 80.50% with 45.50% Mo). Ab initio calculations pro-
vided insights into the intermolecular interactions responsible for
the key features of the novel flotation process. The results revealed
that HS
ions have a higher affinity to desorb collector from
Z200@ccp complex compared with BX@ccp complex. This process
scheme has potential for future separations of other ores with sim-
ilar mineralogical properties.
Declaration of Competing Interest
The authors declare that they have no known competing finan-
cial interests or personal relationships that could have appeared
to influence the work reported in this paper.
Acknowledgments
This work was supported by the financial support from the
National Key R&D Program of China (No. 2022YFC2904502,
No. 2022YFC2904501), Hunan High-tech Industry Technology
Innovation Leading plan (No. 2022GK4056); This work was
supported in part by the High Performance Computing Center of
Central South University.
Appendix A. Supplementary material
Supplementary data to this article can be found online at
https://doi.org/10.1016/j.apt.2023.104039.
References
[1] S.M. Bulatovic, Handbook of flotation reagents: chemistry, theory and practice:
volume 1: flotation of sulfide ores, Elsevier, 2007.
[2] G. Yi, E. Macha, J. Van Dyke, R.E. Macha, T. McKay, M.L. Free, Recent progress on
research of molybdenite flotation: A review, Adv. Colloid Interface Sci. 295
(2021).
[3] S.A. Khoso, Y. Hu, R. Liu, M. Tian, W. Sun, Y. Gao, H. Han, Z. Gao, Selective
depression of pyrite with a novel functionally modified biopolymer in a Cu–Fe
flotation system, Miner. Eng. 135 (2019) 55–63.
[4] X. Zhang, X. Gu, Y. Han, N. Parra-Álvarez, V. Claremboux, S.K. Kawatra,
Flotation of iron ores: a review, Miner. Process. Extr. Metall. Rev. 42 (2021)
184–212.
[5] S. Wu, J. Wang, L. Tao, R. Fan, C. Wang, W. Sun, Z. Gao, Selective separation of
chalcopyrite from pyrite using an acetylacetone-based lime-free process,
Miner. Eng. 182 (2022).
[6] M. Zanin, I. Ametov, S. Grano, L. Zhou, W. Skinner, A study of mechanisms
affecting molybdenite recovery in a bulk copper/molybdenum flotation circuit,
Int. J. Miner. Process. 93 (2009) 256–266.
[7] D. Yuan, K. Cadien, Q. Liu, H. Zeng, Flotation separation of Cu-Mo sulfides by O-
Carboxymethyl chitosan, Miner. Eng. 134 (2019) 202–205.
[8] M. Li, D. Wei, Q. Liu, W. Liu, J. Zheng, H. Sun, Flotation separation of copper–
molybdenum sulfides using chitosan as a selective depressant, Miner. Eng. 83
(2015) 217–222.
[9] B. Yang, H. Yan, M. Zeng, H. Zhu, Tiopronin as a novel copper depressant for the
selective flotation separation of chalcopyrite and molybdenite, Sep. Purif.
Technol. 266 (2021).
[10] A. Ansari, M. Pawlik, Floatability of chalcopyrite and molybdenite in the
presence of lignosulfonates. Part II. Hallimond tube flotation, Miner. Eng. 20
(2007) 609–616.
[11] Z. Yin, W. Sun, Y. Hu, J. Zhai, G. Qingjun, Evaluation of the replacement of NaCN
with depressant mixtures in the separation of copper–molybdenum sulphide
ore by flotation, Sep. Purif. Technol. 173 (2017) 9–16.
[12] J. Chen, L. Lan, X. Liao, Depression effect of pseudo glycolythiourea acid in
flotation separation of copper–molybdenum, Trans. Nonferrous Met. Soc. Chin.
23 (2013) 824–831.
[13] Z. Yin, S. Chen, Z. Xu, C. Zhang, J. He, J. Zou, D. Chen, W. Sun, Flotation
separation of molybdenite from chalcopyrite using an environmentally-
efficient depressant L-cysteine and its adsoption mechanism, Miner. Eng.
156 (2020).
[14] C. Guan, Z. Yin, S. Ahmed Khoso, W. Sun, Y. Hu, Performance analysis of
thiocarbonohydrazide as a novel selective depressant for chalcopyrite in
molybdenite-chalcopyrite separation, Minerals 8 (2018) 142.
[15] Q. Zhang, S. Wen, Q. Feng, S. Zhang, Surface characterization of azurite
modified with sodium sulfide and its response to flotation mechanism, Sep.
Purif. Technol. 242 (2020).
[16] D. Yuan, K. Cadien, Q. Liu, H. Zeng, Adsorption characteristics and mechanisms
of O-Carboxymethyl chitosan on chalcopyrite and molybdenite, J. Colloid
Interface Sci. 552 (2019) 659–670.
[17] C. Wang, R. Liu, M. Wu, Z. Xu, M. Tian, Z. Yin, W. Sun, C. Zhang, Flotation
separation of molybdenite from chalcopyrite using rhodanine-3-acetic acid as
a novel and effective depressant, Miner. Eng. 162 (2021).
[18] T. Hirajima, M. Mori, O. Ichikawa, K. Sasaki, H. Miki, M. Farahat, M. Sawada,
Selective flotation of chalcopyrite and molybdenite with plasma pre-
treatment, Miner. Eng. 66 (2014) 102–111.
[19] Z. Wei, Y. Li, H. Gao, Y. Zhu, G. Qian, J. Yao, New insights into the surface
relaxation and oxidation of chalcopyrite exposed to O
2
and H
2
O: a first-
principles DFT study, Appl. Surf. Sci. 492 (2019) 89–98.
[20] J. Chen, X. Long, Y. Chen, Comparison of multilayer water adsorption on the
hydrophobic galena (PbS) and hydrophilic pyrite (FeS
2
) surfaces: a DFT study,
J. Phys. Chem. C 118 (2014) 11657–11665.
[21] Y. Luo, L. Ou, J. Chen, H. Zhou, C. Yin, H. Yang, Insights into the adsorption
mechanism of water at different coverage rates on talc (Mg
3
Si
4
O
10
(OH)
2
)(00
1) basal surface: a first-principles study, J. Mol. Liq. 363 (2022).
[22] Y. Luo, L. Ou, J. Chen, G. Zhang, Y. Xia, B. Zhu, H. Zhou, Study on the
enhancement mechanism of Pb ion on smithsonite sulfidation with
coordination chemistry and first-principles calculations, Appl. Surf. Sci. 597
(2022).
[23] Y. Luo, L. Ou, G. Zhang, J. Chen, Y. Luo, H. Zhou, H. Yang, C. Yin, Unveiling the
role of Ca ion in the sulfidation of smithsonite: a density functional theory
study, J. Mol. Liq. 367 (2022).
[24] W. Cui, J. Zhang, J. Chen, Surface proximity effect of galena and its influence on
synergistic adsorption behavior, Appl. Surf. Sci. 567 (2021).
[25] Y. Luo, L. Ou, J. Chen, G. Zhang, Y. Xia, B. Zhu, H. Zhou, Insights into the
adsorption performance and mechanism of hydrated Ca ion on talc (001) basal
surface from DFT calculation, Int. J. Min. Sci. Technol. (2022).
[26] J. Liu, Y. Zeng, D. Luo, S. Wen, Y. Wang, Ab initio molecule dynamic simulation
of Cu(OH)
2
interaction with sphalerite (1 1 0) surface, Miner. Eng. 122 (2018)
176–178.
[27] Y. Luo, Y. Xia, H. Zhou, C. Yin, H. Yang, J. Chen, L. Ou, Effect of calcium ions on
surface properties of chalcopyrite and arsenopyrite and its response to
flotation separation under low-alkalinity conditions, Appl. Surf. Sci. (2022).
[28] Q. Feng, S. Wen, J. Deng, W. Zhao, DFT study on the interaction between
hydrogen sulfide ions and cerussite (110) surface, Appl. Surf. Sci. 396 (2017)
920–925.
[29] Y. Luo, L. Ou, J. Chen, G. Zhang, Y. Xia, B. Zhu, H. Zhou, Understanding the
wettability and natural floatability of PbS with different types of vacancy
defects: a perspective from spin-polarized DFT-D and MD, J. Mol. Liq. (2022).
[30] J. Liu, S. Wen, J. Deng, X. Chen, Q. Feng, DFT study of ethyl xanthate interaction
with sphalerite (110) surface in the absence and presence of copper, Appl. Surf.
Sci. 311 (2014) 258–263.
[31] Y. Luo, L. Ou, J. Chen, G. Zhang, K. Jia, Experimental and DFT study of the
sulfidation of smithsonite: impact of water and oxygen, Appl. Surf. Sci. 592
(2022).
[32] M. Tian, C. Zhang, H. Han, R. Liu, Z. Gao, P. Chen, L. Wang, Y. Li, B. Ji, Y. Hu, W.
Sun, Effects of the preassembly of benzohydroxamic acid with Fe (III) ions on
its adsorption on cassiterite surface, Miner. Eng. 127 (2018) 32–41.
[33] Y. Luo, L. Ou, J. Chen, G. Zhang, Y. Xia, B. Zhu, H. Zhou, Mechanism insights into
the hydrated Al ion adsorption on talc (001) basal surface: a DFT study, Surf.
Interfaces (2022).
[34] Y. Luo, L. Ou, J. Chen, G. Zhang, Y. Xia, B. Zhu, H. Zhou, Atomic-level insights
into the modification mechanism of Fe (III) ion on smithsonite (1 0 1) surface
from DFT calculation, Adv. Powder Technol. 33 (2022).
[35] J. Chen, J. Wang, Y. Li, M. Liu, Y. Liu, C. Zhao, W. Cui, Effects of surface spatial
structures and electronic properties of chalcopyrite and pyrite on Z-200
selectivity, Miner. Eng. 163 (2021).
Y. Luo, L. Ou, W. Sun et al. Advanced Powder Technology 34 (2023) 104039
9
Article
In this paper, density functional theory calculations were innovatively used to study the experimentally observed depression of Ca ion and its mechanism on the sulfidation of smithsonite. The calculation results indicated that the reactivity of smithsonite surface after Ca ion adsorption was remarkably decreased, and the adsorption energies for HS on smithsonite surface in the presence of Ca ion were far larger than those in the absence of Ca ion. For one thing, the analysis results of Mulliken bond population, electron density, and partial density of states showed that the generated Ca-S bonds for HS adsorption on smithsonite surface after Ca ion adsorption were weaker and more unstable compared with Zn-S bonds for HS adsorption on the surface before Ca ion adsorption. For another, the hydration shell on smithsonite surface after Ca ion adsorption was found to be more difficult to be broken through by HS. These two factors lead to the inhibition of Ca ion on the adsorption of HS on the surface of smithsonite. This work reveals inhibition of Ca ion and its atomic-level mechanism on the sulfidation of smithsonite, which is helpful to the understanding of the negative effect of Ca ion on the sulfidation of smithsonite.
Article
Study on the interaction between water and mineral surface is significant for the understanding of the subsequent interfacial reactions during the flotation. In this work, the adsorption of water at different coverage rates on talc (0 0 1) basal surface was creatively researched by first-principles calculations. We found the single water adsorption on talc (0 0 1) basal surface was majorly through the H-bond between surface Os and water Hw and H-bond between surface Hs and water Ow with the adsorption energy ranges from -24.26 kJ/mol to -3.8 kJ/mol, and the bonding mechanism of H-bond was attributed to the orbital hybridization of H atom and O atom. Furthermore, we found more H-bonds were formed on talc surface with the increasing water coverage rates, and the average adsorption energy of water was gradually decreased. Thus, the adsorption of water on talc (0 0 1) basal surface was enhanced with the increasing water coverage rates, and the talc surface became more stable after water adsorption. Moreover, charge transfers from water to the talc surfaces occurred, and the work function of talc surfaces gradually decreased when the coverage rates of water were low while increased when the coverage rates of water were higher.
Article
Herein, the effect of calcium ions on surface properties of chalcopyrite and arsenopyrite and its response to flotation separation under low-alkalinity conditions were investigated by flotation tests, zeta potential measurements, solution chemistry analysis, XPS tests, and DFT calculations. The flotation test results showed that calcium ions can selectively decrease the arsenopyrite floatability. The analysis results of zeta potential and solution chemistry showed that the positively charged Ca²⁺ and Ca(OH)⁺ were adsorbed on the chalcopyrite and arsenopyrite surfaces, and more adsorption of which on the arsenopyrite surface was found. Furthermore, the pre-adsorption of calcium ions greatly hindered the collector adsorption on the arsenopyrite surface while did not hinder that on the chalcopyrite surface. The XPS test results further demonstrated that CaOH⁺ was covered on the arsenopyrite surface via interaction with the Fe atom and S atom. Moreover, DFT calculation results revealed that the Ca-S bond can form on the chalcopyrite surface while the Fe-O bond cannot form. However, the Fe-O bond and Ca-S bond were both formed on the arsenopyrite surface, and the formed Ca-S bonds on the chalcopyrite surface were weaker than that on the arsenopyrite surface. Thus, a more hydrophilic surface of arsenopyrite was generated, which decreased its floatability.
Article
Fe(III) ion can strongly inhibit the sulphidation amine flotation of smithsonite. However, its modification mechanism on smithsonite surface is still obscure. In this work, a systematic study of the modification of Fe(III) ion on smithsonite (1 0 1) surface was performed using DFT calculation. The optimal number of H2O ligands for Fe(III) ion hydrates in aqueous conditions was probed, and [Fe(OH)2(H2O)4]⁺ and [Fe(OH)4]⁻ were identified as the major modification species, then their adsorption and bonding mechanisms were further revealed by analyzing the frontier orbitals, density of state, Mulliken population, and electron density. The calculated adsorption structures were consistent with the former experiment, and we found the O site that bonded to the C atom on smithsonite surface was the most favorable position for [Fe(OH)2(H2O)4]⁺ and [Fe(OH)4]⁻ adsorptions. Besides, their adsorption mechanisms on smithsonite surface were principally due to the combined effect of FeO bond and hydrogen bonding. Simultaneously, hydrogen bonding greatly enhanced the stability of the adsorption structures. Moreover, the dominant orbital contribution for the bonding of FeO was primarily due to the orbital hybridization between Fe 3d and O 2p orbitals. This work can help in deeper understanding of the depression of Fe(III) ion on the sulphidation amine flotation of smithsonite.
Article
The utilization of Ca ion as assistant depressant of CMC on talc has been widely reported. Thus, the study on the adsorption mechanism of Ca ion on talc surface is very crucial for understanding the performance of CMC on talc depression. In this paper, mechanism insights into hydrated Ca ion adsorption on talc (0 0 1) basal surface were creatively provided using DFT calculation. [Ca(H2O)6]²⁺ and [Ca(OH)(H2O)3]⁺ were determined as the effective hydrate components for Ca ion adsorption, and the top O site was the most favorable position for their adsorptions on talc surface. Furthermore, the adsorption mechanisms of [Ca(H2O)6]²⁺ and [Ca(OH)(H2O)3]⁺ on talc surface were found to be not the Ca-O chemical bond, but the hydrogen bonding formed by the H atom of the H2O ligand and the surface O atom. H2O acted like a bridge to connect them to the talc surface. Moreover, the hydrogen bonding was formed due to the hybridization of H 1s orbital with the O 2s, O 2p orbitals. Simultaneously, electrons transferred between the H atom and the surface O atom. This work provides theoretical insights into the Ca ion adsorption on talc surface, which can help deeply understand the talc flotation using CMC as depression.
Article
Herein, coordination chemistry analysis and first-principles calculations were creatively employed to investigate the enhancement mechanism of Pb ion on smithsonite sulfidation. The results showed that the reactivity of the Pb ion modified smithsonite surface increased, and HS was found to weakly combine with the low-spin Zn on the smithsonite surface due to the inert Zn 3d orbital while strongly interact with the Pb on Pb modified surface due to the intense ionic polarization interaction caused through the center deviations between the bonding atoms but not the π back-bonding. The Mulliken bond population and PDOS analysis further demonstrated that the bonding of Pb-S was stronger than that of Zn-S. In addition, HS was found to be more likely to break through the hydrated layer on Pb ion modified smithsonite surface. The above two reasons resulted in the enhancement of Pb ion on smithsonite sulfidation. Moreover, analysis results showed that the orbital peaks of smithsonite surface around the Fermi level were depleted after HS adsorption on the Pb ion modified smithsonite surface, and thus a more stable sulfidation product was generated. This work sheds some new light on the enhancement mechanism of Pb ion on smithsonite sulfidation.
Article
Natural galena (PbS) usually contains some defects and impurities, which will greatly affect its wettability and natural floatability. In this paper, the wettability and wetting mechanisms of the perfect PbS (100) surface (PG) and defective surfaces with sulfur vacancy (SV), lead vacancy (PbV), and double neighboring vacancies (DV) were innovatively investigated using spin-polarized Density Functional Theory-Dispersion included (DFT-D) calculations and Molecular Dynamics (MD) simulations. We found the active sites for H2O adsorption on the perfect and defective PbS surfaces were the S atoms, and the reactivity of S at the defect on SV surface decreased, on PbV surface slightly increased while on DV surface greatly increased. Furthermore, we found H-bond between H and S atoms was the main driving force for H2O adsorption, and the single sulfur vacancy increased the hydrophobicity, lead vacancy decreased the hydrophobicity while double neighboring vacancies greatly decreased the hydrophobicity. Thus, the natural floatability of different PbS was in the order of SV>PG>PbV>DV. Moreover, the primary factor to determine the differences in the wettability and natural floatability of different types of PbS was found to be the reactivity of S at the defects but not the steric hindrance effect and the density of the active sites.
Article
In the flotation separation of chalcopyrite and pyrite, the commonly used high-lime process often causes problems such as losses of precious metals, blockage in pipelines and environmental pollution. In this study, acetylacetone (AA) was used as a novel collector to separate chalcopyrite from pyrite without adding limes. The flotation results indicate that AA can selectively separate chalcopyrite from pyrite without depressants at pH 9.0. Flotation kinetics and entrainment calculations show that compared with potassium ethyl xanthate (KEX), AA reduces pyrite entrainment and produces the concentrate with a higher Cu grade and a higher separation efficiency. Solution chemistry calculations show that enolate (AAE⁻) and ketone (AAK) are the dominant species of AA at pH 9.0. Contact angle measurements, zeta potential measurements and Fourier transform infrared (FTIR) analysis indicate that the dominant species of AA chemically adsorb onto chalcopyrite surfaces at pH 9.0 while the adsorption on pyrite is minimal. Density functional theory (DFT) analysis reveals that the Fe atoms on chalcopyrite are the active sites interacting with AA to form six-membered ring complexes. The selective adsorption of AA relies on the spatial distribution and reactivity of Fe atoms on the commonly exposed cleavage surfaces of chalcopyrite and pyrite. After surface relaxation, a layer of Fe atoms appears on the outermost layer of chalcopyrite (1 1 2) surface, which is beneficial for the adsorption of AA species while a layer of S atoms for pyrite (1 0 0) surface. The Fe (III) on chalcopyrite (1 1 2) surface has a higher reactivity than the Fe (II) on pyrite (1 0 0) surface, leading to a higher adsorption energy of AA species on the chalcopyrite. Therefore, this acetylacetone-based lime-free process has a great potential for industrial application in the separation of chalcopyrite from pyrite.
Article
Al ion was reported to have an enhanced depression of Carboxymethyl cellulose (CMC) on talc flotation. However, how Al ion is adsorbed on talc surface remains unclear. Herein, the hydrated Al ion adsorption on talc (001) basal surface was creatively studied by DFT calculations. The effective number of H2O ligands in hydrated Al ion was probed and the optimal adsorption structure was investigated. We found that Al(OH)3(H2O) was the preferred hydrate structure, and the top O site was its optimal adsorption position on talc surface. Furthermore, Al-O chemical bond was not formed, and H2O was found to act as a bridge to connect the Al(OH)3(H2O) to talc surface by H-bond between H of H2O ligand and surface O instead of H-bond between H of Al(OH)3 in Al(OH)3(H2O) and surface O. However, [Al(OH)4]⁻ cannot form the H2O bridge due to the electrostatic repulsion, and its adsorption on talc surface was energetically unsupportive. Moreover, we found that the formation of H-bond was due to the hybridization reaction between O 2p and H 1s orbitals, and slight electron transfers between O and H atoms were also observed. Finally, the Al(OH)3(H2O) was adsorbed on talc surface, which is consistent with the performed experimental result.
Article
Water and oxygen have a significant effect on the sulfidation of smithsonite. However, few studies considered the water and oxygen effects, and the roles of water and oxygen in the sulfidation process remain ambiguous. In this work, the impact of water and oxygen in the sulfidation of smithsonite was creatively investigated by flotation tests, XPS measurements, and DFT calculations. The flotation test results showed that the sulfidation level of smithsonite by directly adding Na2S was not good enough to get a high recovery. The XPS measurement results showed that the sulfidation product of ZnSOn can be formed on smithsonite surface apart from Zn-S under oxygenated aqueous conditions. The DFT simulation results showed that mixed molecular and dissociated adsorption of H2O was observed on smithsonite surface, which prevented the HS adsorption. Besides, it further illustrated the formation of sulfidation products of zinc sulfur oxides in the presence of H2O and O2. Moreover, the sulfidation product of Zn-S can be oxidized by the O2. Thus, it was unstable. In summary, oxygen can be involved in the sulfidation process by dissociated adsorption. This work is crucial and helpful for understanding the sulfidation flotation of smithsonite.