You are on page 1of 6

Minerals Engineering 137 (2019) 118–123

Contents lists available at ScienceDirect

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

Investigation on flotation separation of chalcopyrite from arsenopyrite with T


a novel collector: N-Butoxycarbonyl-O-Isobutyl Thiocarbamate
⁎ ⁎
Jiwei Lu, Zhongyun Tong , Zhitao Yuan , Lixia Li
School of Resources and Civil Engineering, Northeastern University, Shenyang 110819, China

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

Keywords: The efficient separation of chalcopyrite from arsenic-bearing minerals, especially arsenopyrite, remains a
Chalcopyrite challenge in practice. In this work, a novel reagent of N-Butoxycarbonyl-O-Isobutyl Thiocarbamate (NBOIT) was
Arsenopyrite employed in both single and mixed binary mineral flotation, and it proved to be highly effective for the flotation
Flotation separation. Furthermore, the role of the pH and NBOIT concentration was evaluated. Additionally, the me-
Selective adsorption
chanism of the selective separation was investigated systemically via zeta potential measurements, Fourier
XPS
transform infrared (FTIR) analysis and X-ray photoelectron spectroscopy (XPS) analysis. It turns out that the
selective chemisorption of NBOIT on chalcopyrite (in the form of Cu-NBOIT complex) over arsenopyrite is as-
cribed to its the stronger reactivity and the high density of Cu ions on the chalcopyrite surfaces. The XPS analysis
results also confirmed the generation of a Cu-NBOIT complex on the surface of chalcopyrite.

1. Introduction magnesia mixture as a selective depressant to separate arsenopyrite


from a bulk auriferous pyrite concentrate from the Olympias plant in
Arsenic is one of the most dangerous inorganic pollutants, causing Chalkidi. Chen et al. (2001) found that thioglycolic acid (TGA), an
environmental and health emergencies in several areas of the world organic depressant that is known to depress chalcopyrite in sulphide
(Lattanzi et al., 2008). It is a penalty element in many base metal flotation, does not affect the flotation of arsenopyrite. Chanturiya et al.
concentrates that are destined for smelting to produce the relevant (2000) reported that dimethyldithiocarbamate, a coagulant in the flo-
metallic products. It is of economic and environmental interest to re- tation of nonferrous metal ores, significantly depresses arsenopyrite at
move arsenic minerals at an early stage of processing such as flotation. pH 10.3 while chalcopyrite is only affected very weakly. Sirkeci (2000)
Unfortunately, arsenic minerals generally have similar flotation beha- reported similar phenomenon using hexyl thioethylamine as a collector
vior to the valuable minerals with which they are associated. Arseno- in a single mineral study and attributed the differential flotation to the
pyrite (FeAsS), which may associate with chalcopyrite, usually reports higher adsorption rate of hexyl thioethylamine on to chalcopyrite in
to the copper concentrate in conventional flotation due to its similar comparison to arsenopyrite. From a large number of investigations in
flotation properties. Arsenic, in the chalcopyrite concentrate, causes this field, the selective flotation reagents were found to play an im-
environmental problems and also problems in the subsequent pyr- portant role in the collection process.
ometallurgical process (Guo and Yen, 2005; Fornasiero et al., 2001; Alkoxycarbonyl thiocarbamate, which can be made from butyl
Tayebi-Khorami et al., 2017). In industry it has been shown that ar- chloroformate, sodium thiocyanate and isobutanol, has good selectivity
senopyrite is difficult to remove from copper concentrate by conven- for copper, its collecting ability is stronger than Z-200, and its toxicity
tional flotation techniques. and its potential for pollution are lower than for xanthate, while its use
Arsenic removal in sulphide flotation has been studied extensively under low alkaline conditions can effectively lower the amount of lime
with various approaches (Ma and Bruckard, 2009). A range of arsenic used. It can thereby reduce costs and protect the environment of the
minerals including arsenopyrite, tetrahedrite, tennantite and enargite mine, while low alkaline conditions are conducive to the recovery of
occurs in various copper deposits in the world (Plackowski et al., 2012). rare metals and more efficient use of mineral resources. Many sulfide
Abeidu and Almahdy (1980) found that a magnesia mixture forms a ore flotation plants use it instead of xanthate. (Han et al., 2001; Yang
strongly hydrophilic layer on arsenopyrite surfaces which inhibits the and Liu, 2015; Yuan et al., 2007; Zhang et al., 2008). The strong
adsorption of xanthates on arsenopyrite. Matis et al. (1992) used a electron-withdrawing ethoxycorbany group adjacent to C]S group


Corresponding authors.
E-mail addresses: tzhongyun@yeah.net (Z. Tong), yuanzhitao@mail.neu.edu.cn (Z. Yuan).

https://doi.org/10.1016/j.mineng.2019.04.003
Received 29 October 2018; Received in revised form 1 April 2019; Accepted 1 April 2019
0892-6875/ © 2019 Elsevier Ltd. All rights reserved.
J. Lu, et al. Minerals Engineering 137 (2019) 118–123

Table 1 The mineral samples were mixed with 40 mL deionized water and
Chemical compositions of chalcopyrite and arsenopyrite. conditioned for 2 min at a speed of 1920 r/min while the desired pH
Elements Fe S Cu As SiO2 MgO Al2O3 CaO value was adjusted using H2SO4 or NaOH. Lime and NBOIT were added
into the cell one after another with a 3 min conditioning time and fol-
Chalcopyrite 30.4 35.9 33.2 0.1 0.1 0.2 0.1 – lowed by flotation for 5 min. The froth products and tailings were
Arsenopyrite 34.1 18.7 0.2 45.7 0.7 – 0.5 0.1
collected, filtered, and dried. The recovery was calculated based on
weight and elemental contents of the products obtained.
decreases the electron density of the sulfur in N-Butoxycarbonyl-O-
2.3. Zeta potential measurement
Isobutyl Thiocarbamate, which improves NBOIT’s selectivity against
arsenic minerals (Liu et al., 2006). In this study, N-Butoxycarbonyl-O-
The zeta potential values of minerals before and after interacting
Isobutyl Thiocarbamate was employed as a collector to make full use of
with collector were measured by a zeta plus meter (Nano-ZS90 Zeta
its two advantages, i.e., a high selectivity and intense collecting ability,
plus, Malvern Instruments Ltd., UK). 20 mg of −10 μm mineral was
respectively. The flotation performance and selectively of NBOIT on
added to 50 mL of 10−3 mol/L potassium nitrate solution as a sup-
chalcopyrite were evaluated by flotation experiments, and the adsorp-
porting electrolyte and mechanically stirred for 10 min and the pH was
tion mechanism of NBOIT on the mineral surface was also examined by
adjusted to each required value in the range of 3–11 using H2SO4 or
zeta potential measurements, Fourier transform infrared (FTIR) spec-
NaOH. In each case, the zeta potentials were measured and the average
troscopy analysis, and X-ray photoelectron spectroscopy (XPS) analysis.
of three independent measurements was accredited.

2. Materials and methods 2.4. FTIR spectra analysis

2.1. Minerals and reagents The Fourier transform infrared (FTIR) spectra were recorded at a
4 cm−1 resolution in the 4000–400 cm−1 region through KBr disks by a
Both chalcopyrite and arsenopyrite samples were obtained from the Nicolet 380 FTIR spectrometer at room temperature. 1 g of −10 μm
Inner Mongolia Autonomous Region of China. The hand-picked mineral samples was added into 30 mL deionized water with or without NBOIT
samples were crushed and then ground in a porcelain ball mill. Then at the desired pH values and conditioned for 30 min. Afterwards,
the −75 + 38 μm size fraction was obtained by wet-sieving for flota- samples were filtered, purged with deionized water and dried at 35 °C.
tion testing. As shown in Table 1 and Fig. 1, the results of element Approximately 2 mg of the sample was thoroughly mixed with 200 mg
analysis and X-ray diffraction spectra (XRD) of samples indicated that of spectroscopic grade KBr as diluent and then pressed into pellets for
the purity of the samples of chalcopyrite and arsenopyrite was 96.4% FTIR measurements.
and 98.3%, respectively, which met the purity requirement for the
study. Artificial mixed mineral was prepared by mixing chalcopyrite 2.5. XPS analysis
and arsenopyrite samples at a mass ratio of 1:1.
The N-Butoxycarbonyl-O-Isobutyl Thiocarbamate (NBOIT) was in- XPS spectra were recorded using an American Thermo ESCALAB
dustry grade. Its purity was greater than 85% and its chemical formula 250 XI spectrometer using Al Kα X-rays (1486.6 eV) as the sputtering
is CH3CH(CH3)CH2eOe(C]S)eNHe(C]O)eOeC4H9. Other chemi- source at a power of 150 W (15 Kv × 10 Ma). The XPS measurements
cals were of AR grade. Sulfuric acid (H2SO4) and sodium hydroxide were performed inside the analysis chamber operating in a high va-
(NaOH) were used as pH regulators. Deionized water was used cuum of 5.0 × 10−7 Pa. A value of 284.7 eV was adopted as the stan-
throughout all experiments. dard C (1s) to calibrate binding energy.

2.2. Flotation tests 3. Results and discussion

Flotation of single minerals and separation of artificial mixed sam- 3.1. Flotation of single minerals
ples (chalcopyrite and arsenopyrite at a mass ratio 1:1) were carried out
in a XFG mechanical flotation machine with a 40 mL flotation cell. Tests The flotation behaviors of chalcopyrite and arsenopyrite using
with both single and mixed minerals employed a sample mass of 2 g. NBOIT as collector were studied and the results are shown in Fig. 2.
Fig. 2 shows flotation performance of chalcopyrite and arsenopyrite
as a function of NBOIT concentration when the slurry pH was 10. The
results indicate that flotation recovery of chalcopyrite was increased
with increase of NBOIT concentration when it less than 200 mg/L. And
then, flotation recovery of chalcopyrite increased slowly above 93%
with further enhancement of NBOIT concentration. However, in the
whole range of tested collector concentration, the flotation recovery of
arsenopyrite was always below 30%. Therefore, NBOIT had strong
collecting ability and selectivity for chalcopyrite, and good separation
of chalcopyrite from arsenopyrite may be achieved under certain con-
ditions.
Fig. 3 demonstrates the flotation performances of chalcopyrite and
arsenopyrite as a function of slurry pH with NBOIT concentration of
200 mg/L and lime concentration of 7.5 mg/L. The results indicate that
high chalcopyrite recovery was obtained when the slurry pH exceeded
8.5. However, the flotation recovery of chalcopyrite is decreased
sharply when the slurry pH is more alkaline than 10.5. Many studies
have confirmed that oxidation of chalcopyrite occurs and more hy-
droxides appear on its surface (Liu and Laskowski, 1989). The flotation
Fig. 1. X-ray diffraction patterns of chalcopyrite (a) and arsenopyrite (b). recovery of arsenopyrite decreases rapidly from 30.3% to 6.0% with

119
J. Lu, et al. Minerals Engineering 137 (2019) 118–123

Fig. 4. Effect of NBOIT concentration on separation of artificially mixed mi-


Fig. 2. Effect of NBOIT concentration on chalcopyrite and arsenopyrite flota- nerals (pH 10, lime 7.5 mg/L).
tion (pH 10).

Fig. 5. Zeta potentials as a function of pH for chalcopyrite and arsenopyrite


Fig. 3. Effect of slurry pH on flotation of chalcopyrite and arsenopyrite (NBOIT with or without NBOIT (NBOIT 200 mg/L).
200 mg/L).
on chalcopyrite and arsenopyrite before and after the addition of
increase in pH from 4.25 to 6.05. After that, arsenopyrite flotation re- NBOIT. The results are shown in Fig. 5. Chalcopyrite showed a positive
covery decreases slowly with further increasing of pH. Thus, the ap- zeta potential in strong acidic solutions with an IEP of about pH 4.
propriate pH for the separation of chalcopyrite and arsenopyrite is These zeta potential values are consistent with those reported in lit-
8.5–10.5. erature (Fullston et al., 1999; Liu et al., 2017; Feng et al., 2018). In the
presence of collector, zeta potentials of chalcopyrite are significantly
3.2. Separation of artificially mixed minerals shifted to more positive values in the pH range from 3 to 11. This in-
dicated that NBOIT adsorbed on the surface of chalcopyrite. Because
The separation performance of artificially mixed minerals was in- NBOIT is non-ionic, its adsorption could increase the distance of the
vestigated at different NBOIT concentrations. The results are presented shear plane in the electrical double layer from the surface of the mi-
Fig. 4. As the NBOIT concentration increased, the recovery of Cu grows neral, thereby lowering the magnitude of zeta potential. Previous stu-
steadily from 53.9% to 91.2%, while the grade barely decreases from dies showed similar results, in which the adsorption of non-ionic re-
32.0% to 29.6%. The grade and recovery of As has little change and agents decreased the negative zeta potentials of chalcopyrite (Feng
always remained at a low level. With NBOIT concentration of 200 mg/ et al., 2018). However, the zeta potentials of arsenopyrite, and ar-
L, the grade and recovery of Cu could reach 31.5% and 90.3%, re- senopyrite with NBOIT (200 mg/L) did not change. This means that
spectively. The results suggested that an effective separation of chal- NBOIT had little impact on arsenopyrite zeta potentials.
copyrite and arsenopyrite could be achieved with NBOIT collector.
3.4. FTIR spectra analysis
3.3. Zeta potential measurements
FTIR spectroscopy was usually employed to determine the adsorp-
The adsorption of reagents on the minerals surfaces usually affects tion behaviors by analyzing the changes of the characteristic peaks
the surface properties (Xu et al., 2013; Gao et al., 2018; Liu et al., (Zhao et al., 2018; Wu et al., 2018). Therefore, adsorption of NBOIT on
2019a; Liu et al., 2019b). Zeta potential measurements were performed the chalcopyrite and arsenopyrite surfaces was investigated by a FTIR

120
J. Lu, et al. Minerals Engineering 137 (2019) 118–123

Fig. 8. XPS survey spectra of chalcopyrite and chalcopyrite with NBOIT.

Fig. 6. FTIR spectra of chalcopyrite before and after interaction with NBOIT at Fig. 7, the characteristic absorption bands of arsenopyrite are at
pH 10. 1637 cm−1, 1127 cm−1 and 435 cm−1, which agree well with the re-
ported data (Henao et al., 2010; Khummalai and Boonamnuayvitaya,
2005). There are no new pronounced bands for arsenopyrite after
treatment with NBOIT, implying weak adsorption for NBOIT on the
arsenopyrite surface. The weak adsorption is the main reason why
NBOIT barely increases the flotation of arsenopyrite.

3.5. XPS analysis

XPS spectroscopy, as a very surface-sensitive analytical technique,


was widely used to measure the changes occurring at the minerals
surfaces in minerals engineering (Fang et al., 2018; Xu et al., 2017;
Zhang et al., 2018). To further confirm the interaction mechanism be-
tween NBOIT and chalcopyrite, XPS analysis of chalcopyrite and chal-
copyrite treated with NBOIT was conducted, and the results are shown
in Fig. 8. It could be found from Fig. 8 that after NBOIT treatment, the N
1s XPS peak was detected on the chalcopyrite surface. The atomic
concentrations of elements C 1s, O 1s, Cu 2p, Fe 2p, S 2p, and N 1s
measured by XPS were summarized in Table 2. Table 2 shows that the N
1s atomic concentration increased from 3.65% of pure chalcopyrite to
Fig. 7. FTIR spectra of arsenopyrite before and after interaction with NBOIT at 4.04% of chalcopyrite treated with NBOIT, indicating that NBOIT
pH 10. successfully adsorbed on the chalcopyrite surface. Furthermore, the C
1s atomic concentration of chalcopyrite treated with NBOIT increased
2.19%, evidencing that NBOIT adsorbed on to the surfaces of chalco-
spectroscope. FTIR spectra of chalcopyrite and arsenopyrite, NBOIT,
pyrite.
chalcopyrite and arsenopyrite conditioned with NBOIT at pH 10 are
To obtain the detailed information about NBOIT attachment, the
shown in Figs. 6 and 7. In the FTIR spectrum of NBOIT, the bands at
chemical status of surface species was further characterized by high
2961 cm−1 and 2874 cm−1 could be attributed to the eCH3 and eCH2
resolution XPS spectra. The Cu 2p3/2 spectra of chalcopyrite and
stretching vibration. The bands at 1517 cm−1 and 1170 cm−1were at-
chalcopyrite treated with NBOIT are illustrated in Fig. 9. The Cu 2p3/2
tributed to eC(]S)eNHe and eOeC(]O)eNeCe group vibration
spectra are fitted by three components at 932.10 eV, 932.80 eV and
(Leppinen et al., 1988). The bands at around 1064 cm−1 and
934.50 eV, which are assigned to CuFeS2/CuS, CuFeS2/Cu2S and CuO/
1770 cm−1 were assigned to the asymmetric stretching vibration of the
Cu(OH)2, respectively (Fairthorne et al., 1998; Acres et al., 2010). The
C]S and C]O group (Buckley et al., 2014). After interaction with
parameters of the fitted peaks and the percentages of copper occurring
NBOIT, the most intense peaks around 2962 cm−1 and 2871 cm−1 arise
in three surface copper components are shown in Table 3. After NBOIT
from the stretching vibrations ofeCH3 and eCH2 groups in NBOIT
treatment, the proportion of the Cu 2p3/2 band at 934.5 eV on chal-
molecules, indicating the NBOIT adsorption on the chalcopyrite surface
copyrite surface reduced from 3.52% to 2.94%, suggesting that NBOIT
(Huang et al., 2018; Zhang et al., 2013). The bands of 1517 cm−1,
reacted with copper cation and formed new substance.
1170 cm−1 and 1064 cm−1 in the NBOIT FTIR spectrum shifted to
The high resolution spectra of S 2p for samples are displayed in
1507 cm−1, 1173 cm−1 and 1068 cm−1 respectively in the FTIR spec-
Fig. 10. S 2p peaks occur as doublets (S 2p3/2 and S 2p1/2) owing to
trum of chalcopyrite treated with NBOIT. Simultaneously, the char-
spin-orbit splitting (Liu et al., 2015). The S 2p3/2 peaks are twice the
acteristic peak of chalcopyrite around 1436 cm−1 shifted to 1458 cm−1.
area of the S 2p1/2 peak with a 1.19 eV splitting (Yang et al., 2015).
This demonstrates that the adsorption of NBOIT on the surface of
The parameters of the fitted peaks and the corresponding assigned
chalcopyrite is by the chemisorption process. Thus, the chemisorption
species were obtained from the reported data and are shown in Table 4.
may have occurred between NBOIT and the chalcopyrite surface. From
The most intense doublet centered at 161.15/162.33 eV is related to the

121
J. Lu, et al. Minerals Engineering 137 (2019) 118–123

Table 2
Atomic concentration of elements on chalcopyrite surfaces as determined by XPS.
Samples Atomic concentration of elements on chalcopyrite surfaces (%)

C 1s O 1s Cu 2p Fe 2p S 2p N 1s

Chalcopyrite 16.06 37.07 11.36 12.48 19.38 3.65


Chalcopyrite + NBOIT 18.25 28.31 14.61 13.95 20.84 4.04

Table 4
The parameters of fitted peaks and percentages of surface sulfur components in
total sulfur on the samples.
Binding energy (eV) 161.15/ 162.11/ 163.10/ 164.03/
162.33 163.29 164.22 165.21

FWHM (eV) 0.88/0.88 0.88/0.88 0.88/0.88 0.88/0.88


Assignment S2− S2- n S2- 2 Organic S
Chalcopyrite 80.92 14.01 5.07 –
Treated chalcopyrite 1.42 79.05 17.70 1.83

monosulfides (S2−) in CuFeS2. The other two weak doublets positioned


at 162.11/163.29 eV and 163.10/164.22 eV could be attributed to
oxidized sulfur components (polysulfides S2− n and disulfides S2− 2)
(Acres et al., 2010; Harmer et al., 2004; Harmer et al., 2006). After
NBOIT adsorption, the S 2p3/2 spectra may be divided into four com-
ponents, i.e. S2− (161.15 eV), S2- n(162.11 eV), S2- 2(163.10 eV) and
organic S (164.03 eV). The corresponding proportion of these various
species is 1.42%, 79.05%, 17.70% and 1.83%. In the presence of ad-
Fig. 9. High resolution XPS spectra of Cu 2p3/2 ((a) chalcopyrite and (b) sorbed NBOIT only a small part of S2− on the chalcopyrite surface was
chalcopyrite treated with NBOIT). detected. The increase of S2− 2 species at around 163.10 eV from
5.07% to 17.70% (by 12.63%) was far less than that at around
162.11 eV from 14.01% to 79.05% (by 65.04%). This finding implies a
Table 3
The parameters of fitted peaks and percentage of surface copper components in
decreased oxidation of chalcopyrite surfaces after NBOIT coverage. The
total copper on the samples. higher binding energy component at 164.03 eV might represent organic
sulfur in the Cu-NBOIT complex. The strong electron-withdrawing
Binding energy (eV) 932.1 932.8 934.5
ethoxycarbonyl group adjacent to the C]S group decreases the electron
FWHM (eV) 1.1 1.74 1.61 density of sulfur, which improves NBOIT’s selectivity at alkaline con-
Assignment CuFeS2/CuS CuFeS2/Cu2S CuO/Cu(OH)2 ditions. NBOIT’s LUMO (Lowest Unoccupied Molecular Orbital) is
Chalcopyrite 89.57 6.91 3.52 constituted by pz-orbits of every atom in the conjugate eOeC(]O)
Treated chalcopyrite 92.6 4.46 2.94
eNeC(]S)eOe group and willingly accepts strongly d-orbital elec-
trons from Cu(II) or Cu(I) configurations of the copper cation on the
chalcopyrite surface. Hence, the interaction of NBOIT with the copper
cation involves forming a six-membered chelate ring, and the collecting
power of NBOIT for chalcopyrite and its selectivity are improved at the
same time (Liu et al., 2006; Shen et al., 1998). Based on the results and
analyses mentioned above, a schematic adsorption model of NBOIT on
chalcopyrite surfaes was suggested in Fig. 11.

4. Conclusion

In the present work, we found that a novel reagent N-


Butoxycarbonyl-O-Isobutyl Thiocarbamate (NBOIT) as a collector along

Fig. 10. High resolution XPS spectra of S 2p spectra ((a) chalcopyrite and (b)
chalcopyrite treated with NBOIT).

Fig. 11. The suggested chemisorption model of NBOIT on chalcopyrite surfaces.

122
J. Lu, et al. Minerals Engineering 137 (2019) 118–123

with lime as a depressor, can achieve the selective separation of chal- layers formed during chalcopyrite leaching. Geochim. Cosmochim. Acta. 70,
copyrite from arsenopyrite in the flotation of mixed binary minerals at 4392–4402. https://doi.org/10.1016/j.gca.2006.06.1555.
Henao, D.M.O., Godoy, M.A.M., 2010. Jarosite pseudomorph formation from arseno-
pH 10. Diverse analytical methods, including zeta potential measure- pyrite oxidation using Acidithiobacillus ferrooxidans. Hydrometallurgy 104,
ments, FTIR spectroscopy and XPS analysis were employed to elucidate 162–168. https://doi.org/10.1016/j.hydromet.2010.05.012.
the adsorption mechanism. Based on the analysis, we draw the fol- Huang, Y.G., Liu, G.Y., Ma, L.Q., Liu, J., 2018. 5-Heptyl-1,3,4-oxadiazole-2-thione:
synthesis and flotation mechanism to chalcopyrite. J. Ind. Eng. Chem. 61, 331–339.
lowing conclusions: https://doi.org/10.1016/j.jiec.2017.12.031.
Khummalai, N., Boonamnuayvitaya, V., 2005. Suppression of arsenopyrite surface oxi-
(i) NBOIT has strong collecting ability and selectivity for chalcopyrite. dation by sol-gel coatings. J. Biosci. Bioeng. 99, 277–284. https://doi.org/10.1263/
jbb.99.277.
(ii) Selective chemisorption of NBOIT on to chalcopyrite occurs in the Lattanzi, P., Da, P.S., Musu, E., Atzei, D., Elsener, B., Fantauzzi, M., Rossi, A., 2008.
form of a Cu-NBOIT complex on the surface of chalcopyrite. Enargite oxidation: a review. Earth-Sci. Rev. 86, 62–68. https://doi.org/10.1016/j.
earscirev.2007.07.006.
Leppinen, J.O., Basilio, C.I., Yoon, R.H., 1988. FTIR study of thionocarbamate adsorption
In summary, NBOIT can selectively adsorb on the chalcopyrite
on sulfide minerals. Coll. Surf. A-Physicochem. Eng. Asp. 32, 113–125. https://doi.
surface involving intense chemisorption, while NBOIT exhibits weak org/10.1016/0166-6622(88)80008-8.
adsorption to arsenopyrite. The adsorption can promote the hydro- Liu, C., Zhang, W.C., Song, S.X., Li, H.Q., 2019a. a. A novel method to improve carbox-
phobicity of the chalcopyrite surface, which is very important for a ymethyl cellulose performance in the flotation of talc. Miner. Eng. 131, 23–27.
https://doi.org/10.1016/j.mineng.2018.11.003.
successful separation. It has great potential to be applied industrially to Liu, C., Zhang, W.C., Song, S.X., Li, H.Q., Jiao, X.K., 2019b. b. A novel insight of the effect
deal with abundant complicated chalcopyrite bearing ore, especially of sodium chloride on the sulfidization flotation of cerussite. Powder Technol. 344,
those associated with arsenopyrite. 103–107. https://doi.org/10.1016/j.powtec.2018.12.002.
Liu, G.Y., Qiu, Z.H., Wang, J.Y., Liu, Q.X., Xiao, J.J., Zeng, H.P., Zhong, H., Xu, Z.H.,
2015. Study of N-isopropoxypropyl-N’-ethoxycarbonyl thiourea adsorption on chal-
Acknowledgments copyrite using in situ SECM, ToF-SIMS and XPS. J. Coll. Interf. Sci. 437, 42–49.
https://doi.org/10.1016/j.jcis.2014.08.069.
Liu, G.Y., Zhong, H., Dai, T.G., 2006. The separation of Cu/Fe sulfide minerals at slightly
This work was supported by the National Natural Science alkaline conditions by using ethoxycarbonyl thionocarbamates as collectors: theory
Foundation of China (51704057); the Fundamental Research Funds for and practice. Miner. Eng. 19, 1380–1384. https://doi.org/10.1016/j.mineng.2005.
the Central Universities (N170104018); the China Postdoctoral Science 12.007.
Liu, Q., Laskowski, J.S., 1989. The role of metal hydroxides at mineral surfaces in dextrin
Foundation (2017 M621153); and the Postdoctoral Science Foundation adsorption, II. chalcopyrite-galena separations in the presence of dextrin. Int. J.
of Northeastern University (20170312). Miner. Process. 27, 147–155. https://doi.org/10.1016/0301-7516(89)90012-4.
Liu, S., Liu, G.Y., Zhong, H., Yang, X.L., 2017. The role of HABTC’s hydroxamate and
dithiocarbamate groups in chalcopyrite flotation. J. Ind. Eng. Chem. 52, 359–368.
Appendix A. Supplementary material
https://doi.org/10.1016/j.jiec.2017.04.015.
Ma, X., Bruckard, W.J., 2009. Rejection of arsenic minerals in sulfide flotation – a lit-
Supplementary data to this article can be found online at https:// erature review. Int. J. Miner. Process. 93, 89–94. https://doi.org/10.1016/j.minpro.
doi.org/10.1016/j.mineng.2019.04.003. 2009.07.003.
Matis, K.A., Kydros, K.A., Gallios, G.P., 1992. Processing a bulk pyrite concentrate by
flotation reagents. Miner. Eng. 5, 331–342. https://doi.org/10.1016/0892-6875(92)
References 90215-U.
Plackowski, C., Nguyen, A.V., Bruckard, W.J., 2012. A critical review of surface prop-
erties and selective flotation of enargite in sulphide systems. Miner. Eng. 30, 1–11.
Abeidu, A.M., Almahdy, A.M., 1980. Magnesia mixture as a regulator in the separation of https://doi.org/10.1016/j.mineng.2012.01.014.
pyrite from chalcopyrite and arsenopyrite. Int. J. Miner. Process 6, 285–302. https:// Shen, I.W., Fornasiero, D., Ralston, J., 1998. Effect of collectors, conditioning pH and
doi.org/10.1016/0301-7516(80)90026-5. gases in the separation of sphalerite from pyrite. Miner. Eng. 11, 145–158. https://
Acres, R.G., Harmer, S.L., Beattie, D.A., 2010. Synchrotron XPS, NEXAFS, and ToF-SIMS doi.org/10.1016/S0892-6875(97)00147-7.
studies of solution exposed chalcopyrite and heterogeneous chalcopyrite with pyrite. Sirkeci, A.A., 2000. The flotation separation of pyrite from arsenopyrite using hexyl
Miner. Eng. 23, 928–936. https://doi.org/10.1016/j.mineng.2010.03.007. thioethylamine as collector. Int. J. Miner. Process. 60, 263–276. https://doi.org/10.
Buckley, A.N., Hope, G.A., Lee, K.C., Petrovic, E.A., Woods, R., 2014. Adsorption of O- 1016/S0301-7516(00)00023-5.
isopropyl-N-ethyl thionocarbamate on Cu sulfide ore minerals. Miner. Eng. 69, Tayebi-Khorami, M., Manlapig, E., Forbes, E., Bradshaw, D., Edraki, M., 2017. Selective
120–132. https://doi.org/10.1016/j.mineng.2014.08.002. flotation of enargite from copper sulphides in Tampakan deposit. Miner. Eng. 112,
Chanturiya, V.A., Nedosekina, T.V., Fedorov, A.A., 2000. Several features of interaction 1–10. https://doi.org/10.1016/j.mineng.2017.06.021.
between sulfhydryl absorbents of xanthogenate class and dithiocarbamates with Wu, H.Q., Tian, J., Xu, L.H., Fang, S., Zhang, Z.Y., Chi, R., 2018. Flotation and adsorption
pyrite and arsenopyrite. Tsvet. Met. 5, 12–15. of a new mixed anionic/cationic collector in the spodumene-feldspar system. Miner.
Chen, J.H., Feng, Q.M., Lu, Y.P., 2001. Electrochemical mechanism of thioglycolic acid Eng. 127, 42–47. https://doi.org/10.1016/j.mineng.2018.07.024.
depression sulfide minerals. Trans. Nonferr. Met. Soc. China. 11, 145–149. Xu, L.H., Tian, J., Wu, H.Q., Deng, W., Yang, Y.H., Sun, W., Gao, Z.Y., Hu, Y.H., 2017.
Fang, S., Xu, L.H., Wu, H.Q., Tian, J., Lu, Z.Y., Sun, W., Hu, Y.H., 2018. Adsorption of Pb New insights into the oleate flotation response of feldspar particles of different sizes:
(II)/benzohydroxamic acid collector complexes for ilmenite flotation. Miner. Eng. anisotropic adsorption model. J. Coll. Interf. Sci. 505, 500–508. https://doi.org/10.
126, 16–23. https://doi.org/10.1016/j.mineng.2018.06.022. 1016/j.jcis.2017.06.009.
Fairthorne, G., Brinen, J.S., Fornasiero, D., Nagaraj, D.R., Ralston, J., 1998. Spectroscopic Xu, L.H., Wu, H.Q., Dong, F.Q., Wang, L., Wang, Z., Xiao, J.H., 2013. Flotation and ad-
and electrokinetic study of the adsorption of butyl ethoxycarbonyl thiourea on sorption of mixed cationic/anionic collectors on muscovite mica. Miner. Eng. 41,
chalcopyrite. Int. J. Miner. Process. 54, 147–163. https://doi.org/10.1016/S0301- 41–45. https://doi.org/10.1016/j.mineng.2012.10.015.
7516(98)00019-2. Yang, X.J., Liu, Y., 2015. Preparation of N-Butoxycarbonyl-O-Isobutyl thiocarbamate and
Feng, B., Peng, J.X., Zhang, W.P., Ning, X.H., Guo, Y.T., Zhang, W.Z., 2018. Use of locust its flotation performance. Liaoning Chem. Industry 44, 1080–1082.
bean gum in flotation separation of chalcopyrite and talc. Miner. Eng. 122, 79–83. Yang, Y., Harmer, S., Chen, M., 2015. Synchrotron-based XPS and NEXAFS study of
https://doi.org/10.1016/j.mineng.2018.03.044. surface chemical species during electrochemical oxidation of chalcopyrite.
Fornasiero, D., Fullston, D., Li, C., Ralston, J., 2001. Separation of enargite and tennantite Hydrometallurgy 156, 89–98. https://doi.org/10.1016/j.hydromet.2015.05.011.
from non-arsenic copper sulfide minerals by selective oxidation or dissolution. Int. J. Yuan, L., Zhong, H., Liu, G.Y., 2007. Preparation and application of isothiocyanate. Fine
Miner. Process 61, 109–119. https://doi.org/10.1016/S0301-7516(00)00029-6. Chem. Intermed. 37, 10–13.
Fullston, D., Fornasiero, D., Ralston, J., 1999. Zeta potential study of the oxidation of Zhang, W.C., Honaker, R.Q., 2018. Flotation of monazite in the presence of calcite part II:
copper sulfide minerals. Colloid Surf. A-Physicochem. Eng. Asp. 146, 113–121. enhanced separation performance using sodium silicate and EDTA. Miner. Eng. 127,
https://doi.org/10.1016/S0927-7757(98)00725-0. 318–328. https://doi.org/10.1016/j.mineng.2018.01.042.
Gao, Y.S., Gao, Z.Y., Sun, W., Yin, Z.G., Wang, J.J., Hu, Y.H., 2018. Adsorption of a novel Zhang, X.P., Jing, H.Z., Miao, J.H., Lu, S.W., 2008. Progress in the synthesis of thio-
reagent scheme on scheelite and calcite causing an effective flotation separation. J. carbamates. Prog. Chem. 20, 1102–1107.
Coll. Interf. Sci. 512, 39–46. https://doi.org/10.1016/j.jcis.2017.10.045. Zhang, Y.H., Cao, Z., Cao, Y.D., Sun, C.Y., 2013. FTIR studies of xanthate adsorption on
Guo, H., Yen, W.T., 2005. Selective flotation of enargite from chalcopyrite by electro- chalcopyrite, pentlandite and pyrite surfaces. J. Mol. Struct. 1048, 434–440. https://
chemical control. Miner. Eng. 18, 605–612. https://doi.org/10.1016/j.mineng.2004. doi.org/10.1016/j.molstruc.2013.06.015.
10.005. Zhao, L., Liu, W.G., Duan, H., Yang, T., Li, Z., Zhou, S.J., 2018. Sodium carbonate effects
Han, E.S., Luan, R., Gao, C.H., 2001. Research progress of phase transfer catalysts in on the flotation separation of smithsonite from quartz using N, N′-dilauroyl ethyle-
organic synthesis. J. Heibei University Technol. 30, 89–95. nediamine dipropionate as a collector. Miner. Eng. 126 (2018), 1–8. https://doi.org/
Harmer, S.L., Pratt, A.R., Nesbitt, W.H., Fleet, M.E., 2004. Sulfur species at chalcopyrite 10.1016/j.mineng.2018.06.020.
(CuFeS2) fracture surfaces. Am. Miner. 89, 1026–1032.
Harmer, S.L., Thomas, J.E., Fornasiero, D., Gerson, A.R., 2006. The evolution of surface

123

You might also like