Qiancheng Zhang, Limin Zhang, Feng Jiang, Honghu Tang, Li Wang, and Wei Sun, Ferric ion-triggered surface oxidation of galena for efficient chalcopyrite–galena separation, Int. J. Miner. Metall. Mater., 31(2024), No. 2, pp. 261-267. https://doi.org/10.1007/s12613-023-2674-x
Cite this article as:
Qiancheng Zhang, Limin Zhang, Feng Jiang, Honghu Tang, Li Wang, and Wei Sun, Ferric ion-triggered surface oxidation of galena for efficient chalcopyrite–galena separation, Int. J. Miner. Metall. Mater., 31(2024), No. 2, pp. 261-267. https://doi.org/10.1007/s12613-023-2674-x
Research Article

Ferric ion-triggered surface oxidation of galena for efficient chalcopyrite–galena separation

+ Author Affiliations
  • Corresponding author:

    Feng Jiang    E-mail: feng_jiang@csu.edu.cn

  • Received: 15 March 2023Revised: 8 May 2023Accepted: 11 May 2023Available online: 12 May 2023
  • The efficient separation of chalcopyrite (CuFeS2) and galena (PbS) is essential for optimal resource utilization. However, finding a selective depressant that is environmentally friendly and cost effective remains a challenge. Through various techniques, such as microflotation tests, Fourier transform infrared spectroscopy, scanning electron microscopy (SEM) observation, X-ray photoelectron spectroscopy (XPS), and Raman spectroscopy measurements, this study explored the use of ferric ions (Fe3+) as a selective depressant for galena. The results of flotation tests revealed the impressive selective inhibition capabilities of Fe3+ when used alone. Surface analysis showed that Fe3+ significantly reduced the adsorption of isopropyl ethyl thionocarbamate (IPETC) on the galena surface while having a minimal impact on chalcopyrite. Further analysis using SEM, XPS, and Raman spectra revealed that Fe3+ can oxidize lead sulfide to form compact lead sulfate nanoparticles on the galena surface, effectively depressing IPETC adsorption and increasing surface hydrophilicity. These findings provide a promising solution for the efficient and environmentally responsible separation of chalcopyrite and galena.
  • Copper and lead are crucial nonferrous metals with an important role in the growth of the national economy [12]. These metals are commonly extracted from sulfide ores in the form of chalcopyrite (CuFeS2) and galena (PbS). However, chalcopyrite and galena often occur closely together [3], making their efficient separation crucial for optimal resource utilization. In this regard, developing selective and effective depressants for chalcopyrite–galena separation is a top priority in the field of mineral engineering [4].

    Given the low floatability of galena, most chalcopyrite–galena separation methods involve galena inhibition. Potassium dichromate (K2Cr2O7), the most commonly used depressant, reacts with galena to form hydrophilic lead chromate on the surfaces of galena [5]. However, the release of chromium into the water makes dichromate environmentally harmful and unsuitable for long-term use [6]. Organic depressants, including dextrin [7], O,O-bis (2,3-dihydroxypropyl) dithiophosphate (DHDTP) [8], sodium pyrophosphate (SPH) and sodium citrate (SCT) [9], sodium humate [10], and high-molecular polymers [1116], have gained attention as environmentally friendly alternatives to potassium dichromate. Despite having some advantages, organic depressants still face challenges, such as difficulty in degradation, high cost, and poor solubility, limiting their potential for large-scale industrial applications. As a result, the efficient and environmentally responsible separation of chalcopyrite and galena remains a challenge.

    Surface oxidation has been commonly used to reduce the floatability of certain sulfide minerals by generating hydrophilic oxide species on surfaces [1718]. Hydrogen peroxide (H2O2) has been found to be an effective oxidant for galena while having minimal impact on chalcopyrite flotation [1922]. However, the poor stability of strong oxidants remains an issue. Ferric (Fe3+) ions, which have been overlooked due to their relatively weak oxidation capacity, may provide a solution. In our previous work, we discovered that Fe3+ could oxidize molybdenite [23], thus offering a new perspective on the use of cheap, stable, and environmentally friendly Fe3+ as a selective depressant for chalcopyrite–galena separation.

    The effect of Fe3+ on the flotation of chalcopyrite and galena was systematically investigated in this study through a series of single-mineral flotation tests. The results showed that Fe3+ possesses high capability for the selective inhibition of galena. Further investigation utilizing Fourier transform infrared spectroscopy (FT-IR), X-ray photoelectron spectroscopy (XPS), Raman spectroscopy, and scanning electron microscope (SEM) observation revealed that the inhibition mechanism of Fe3+ occurred through the formation of lead sulfate (PbSO4) nanoparticles on the galena surface after oxidation. This study clearly illustrates the capability of Fe3+ for the strong and selective inhibition of galena flotation and presents a novel and promising approach for the separation of chalcopyrite and galena.

    The pure bulk chalcopyrite used in this study was purchased from Jiangxi, China, and pure bulk galena originated from Inner Mongolia, China. The minerals underwent a comprehensive preparation process consisting of three stages. In the first stage, the minerals were crushed by using a jaw crusher, with the size of the particles being reduced to −2 mm and impurities being removed via handpicking. In the second stage, the crushed samples underwent dry ball milling to a particle size of −74 μm. In the third stage, the fraction of the sample within the size range of 38–74 μm was screened out for microflotation tests, whereas the remaining fraction (−38 μm) was reserved for subsequent analyses (XRD, FT-IR, XPS, Raman, and SEM).

    Isopropyl ethyl thionocarbamate (IPETC) with a purity of 95wt% was purchased from MACKLIN, China. Analytical-grade terpineol and FeCl3·6H2O were procured from Sinopharm Chemical Reagent Co., Ltd., China. Deionized water (DI) was used for all experiments.

    Microflotation experiments were conducted by using an XFG flotation machine manufactured by Jilin Mining Machinery Manufacturing Co., Ltd., equipped with a 40 mL Plexiglass cell. For each test, a mixture of 2 g of chalcopyrite or galena sample and 40 mL of DI was added to the cell. Next, a solution of Fe3+ was added to the slurry and stirred for 2 min. Subsequently, IPETC was added as the collector. Then, terpineol was added as the frother. After allowing the slurry to stabilize for 30 s, the foam layer was scraped evenly by hand for 3 min. The flotation products were then filtered, dried, and weighed to calculate flotation recovery.

    To determine the composition of the chalcopyrite and galena samples, XRD analysis was performed by using a Bruker D8 diffractometer equipped with monochromatic Cu Kα radiation and a wavelength of 1.5406 Å. As shown in Fig. 1, all peaks in the spectra matched the standard crystal planes of chalcopyrite (PDF#37-0471) and galena (PDF#05-0592), demonstrating the high degree of the purity of the prepared samples.

    Fig. 1.  XRD patterns of (a) chalcopyrite and (b) galena samples.

    The adsorption of the collector onto the mineral surface was characterized by using FT-IR spectroscopy (Nicolet AVATAR 360 FTIR spectrometer) within the wavenumber range of 4000 to 400 cm−1. Infrared spectra were recorded by applying the KBr tableting technique in transmission mode. The sample mineral was mixed with the desired reagent solution for 30 min. Then, the resulting slurry was filtered and washed with DI three times. The resulting solid was then dried in a vacuum drying oven at 45°C to obtain the test sample.

    The surface valence state was analyzed by using XPS (ESCALAB 250Xi, Thermo Fisher Scientific) and Raman spectroscopy (Renishaw InVia). XPS spectra were obtained within the energy range of 0–1200 eV. For Raman spectroscopy measurements, samples were placed on a glass slide, and measurements were performed at a power of 25 mW in the shift range of 200–1200 cm−1. The morphology and composition of the surface were examined by using field emission SEM (FE-SEM, JSM-7610FPlus, JEOL) equipped with an energy dispersive spectroscopy (EDS) device (ULTIM MAX 40, Oxford). The samples were affixed to small glassware by using conductive glue, then placed in a vacuum and imaged at various magnifications to locate the target. SEM images were generated when all analyses were completed. The test samples for XPS, Raman, and SEM analyses were the same as those for FT-IR measurements.

    Ferrous ions (Fe2+) were detected by measuring the standard titration solution of potassium dichromate with sodium diphenylaniline sulfonate as an indicator. After filtering the reaction solution, the resulting filtrate was used as the detection sample for Fe2+. Typically, a mixture of 2 mL of the sample, 10 mL of sulfur phosphorus mixed acid, and 4–5 drops of sodium diphenylaniline sulfonate was injected into 100 mL of DI and titrated immediately with the standard titration solution of potassium dichromate until the solution turned purple. The concentration of Fe2+ was then obtained by reading and calculating the titration value.

    IPETC is one of the most commonly adopted selective collectors for chalcopyrite. The suitable dosage of IPETC was first explored by adding 1.5 × 10−4 mol/L terpineol as a frother for chalcopyrite and galena. Fig. 2(a) shows that the recoveries of chalcopyrite and galena gradually increased with the increase in IPETC dosage. When the IPETC dosage reached 50 mg/L, approximately 90% of chalcopyrite and 80% of galena could be collected. Therefore, 50 mg/L was selected as the optimal amount of collector for further experiments. In addition, the flotation results showed the difficulty of the efficient separation of chalcopyrite and galena by using IPETC alone despite the high selectivity of IPETC. Fig. 2(b) depicts the effects of the presence of Fe3+ on the floatability of chalcopyrite and galena. The recovery of galena dropped sharply (above 60%) when the concentration of Fe3+ was increased from 0 to 25 mg/L and finally reached zero when 50 mg/L of Fe3+ was added. By contrast, Fe3+ had a negligible influence on chalcopyrite flotation, and chalcopyrite recovery remained at 80% in the presence of Fe3+. The flotation results clearly indicate the selective inhibition ability of Fe3+ for galena flotation.

    Fig. 2.  Flotation recoveries of chalcopyrite and galena as functions of (a) IPETC concentration (1.5 × 10−4 mol/L terpilenol) and (b) Fe3+ concentration (50 mg/L IPETC and 1.5 × 10−4 mol/L terpilenol).

    The adsorption of IPETC onto mineral surfaces was investigated by using the FT-IR technique. Fig. 3(a) shows the FT-IR spectrum of IPETC, in which the peaks at 1053 and 1100 cm−1 are attributed to the characteristic group C=S of IPETC [2426]. The peak located at 1213 cm−1 corresponds to the C−O−C group [2728], whereas those at 1452 and 1522 cm−1 are ascribed to the −CH2 and −CH3 groups [29], respectively. As shown in Fig. 3(b), the characteristic peaks of the C=S and C−O−C groups were detected on the chalcopyrite surface after IPETC interaction and maintained high intensities even with previous Fe3+ treatment, indicating the strong chemosorption of IPETC onto the chalcopyrite surface. Fig. 3(c) shows that IPETC also exhibited obvious adsorption onto the pristine galena surface but rarely adsorbed onto Fe3+-treated surfaces, as evidenced by the disappearance of the peaks of the C=S and C−O−C groups. The findings of FT-IR studies agree with the flotation results and verify that Fe3+ selectively hindered the chemosorption of IPETC onto the galena surface to depress the floatability of the galena.

    Fig. 3.  FT-IR spectra of (a) IPETC, (b) chalcopyrite, and (c) galena in various agent systems.

    The surface morphologies of minerals were analyzed by using SEM to examine the impact of Fe3+ on galena. As illustrated in Fig. 4, the initial chalcopyrite and galena surfaces appeared relatively smooth with visible fragments. After treatment with Fe3+, the chalcopyrite surface retained its original morphology (Fig. 4(b)), whereas the surface of galena became covered with a thin layer of fine particles (Fig. 4(d)). EDS analysis revealed a significant increase in the amount of oxygen on the galena surface (from 10at% to 39.81at%) after reaction with Fe3+, indicating the oxidation of the galena surface. Indubitably, the formation of oxide species on the galena surface can dramatically increase hydrophilicity and obstruct chemosorption with the collector [30]. Notably, no iron elements were detected on the galena surface after treatment with Fe3+, indicating the absence of iron species adsorption.

    Fig. 4.  SEM images of (a) pristine chalcopyrite, (b) Fe3+-treated chalcopyrite, (c) pristine galena, and (d) Fe3+-treated galena; EDS spectra of (e) pristine galena and (f) Fe3+-treated galena.

    Further XPS tests were conducted to identify the type of oxide species present on galena surfaces through the analysis of elemental compositions and bonding states. As shown in Fig. 5, the signals of Pb, S, O, and C elements were detected in the full survey spectra of galena. C was added artificially to the sample during tests to calibrate the peak binding energy [31]. The presence of O can be attributed to the air adsorption or surface oxidation of galena samples [32].

    Fig. 5.  Full survey XPS spectra of original and Fe3+-treated galena samples.

    The high-resolution Pb 4f spectra of pristine galena can be split into four peaks (Fig. 6(a)). The peaks located at 137.6 and 142.5 eV are assigned to the Pb 4f7/2 and Pb 4f5/2 spin-orbit doublets of Pb(II)−S [3334], whereas those at 138.1 and 143.0 eV are attributed to the trace amount of lead oxide (Pb(II)−O) species on the mineral surface. After Fe3+ treatment, two new peaks at 139.9 and 144.7 eV, which are consistent with the Pb 4f peaks of PbSO4, were observed, indicating the presence of PbSO4 on the galena surface after the reaction with Fe3+ [35]. In the S 2p spectra (Fig. 6(b)), the peaks located at 160.9 and 162.1 eV are assigned to the S 2p3/2, and S 2p1/2 spin-orbit doublets of S2− in PbS and those at 168.6 and 169.8 eV are ascribed to PbSO4 [3436]. Fe3+ treatment caused a noticeable increase in the formation of PbSO4, as evidenced by the increased intensity of the S 2p peaks of PbSO4.

    Fig. 6.  (a) Pb 4f and (b) S 2p high-resolution XPS spectra of galena.

    The composition of the chalcopyrite and galena samples was further characterized by using Raman spectroscopy. Fig. 7(a) presents the Raman spectra of chalcopyrite, wherein the dominant peak at approximately ~292 cm−1 is attributed to the A1 mode of the two pairs of S2− vibration in the CuFeS2 lattice [3738]. The two weak peaks at 320 and 353 cm−1 correspond to the vibrational modes associated with S2−, Cu+, and Fe3+ (B2/E modes). Fe3+ treatment did not result in any noticeable changes, demonstrating the high stability of chalcopyrite in Fe3+ solutions. As shown in Fig. 7(b), two new peaks at 446 and 974 cm−1, which are associated with the v1 and v2 modes of the stretching vibration of ${\mathrm{SO}}_4^{2-} $ bands, respectively, appeared in the Raman spectrum of galena after the reaction with Fe3+, further confirming the formation of PbSO4 on the galena surface [3942].

    Fig. 7.  Raman spectra of (a) chalcopyrite and (b) galena.

    Fe3+ ions are commonly present in flotation systems because they are released from the steel balls used for grinding. Previous studies have reported that the produced iron oxidation species plays a dominant role in depressing mineral floatability by adsorbing on mineral surfaces [4344]. However, in this study, Fe3+ was found to inhibit galena efficiently and selectively via surface oxidation instead of iron oxidation species adsorption. Fig. 8 shows that the concentration of Fe2+ gradually increased after Fe3+ was added to the galena pulp. This increase further reflected the redox reaction between Fe3+ and galena (Eq. (1)). The galena inhibition mechanism of Fe3+ is clearly illustrated in Fig. 9 on the basis of all of the analyses conducted in this work. First, the added Fe3+ can quickly react with PbS and induce the formation of PbSO4 nanoparticles on the galena surface. Second, as the reaction proceeds, the PbSO4 nanoparticles form a compact layer on the galena surface, effectively depressing collector adsorption and increasing surface hydrophilicity. Third, Fe3+ exerts a negligible influence on chalcopyrite, thus realizing the selective floating of chalcopyrite.

    Fig. 8.  Variation in the concentration of Fe2+ in the galena system (with the addition of 50 mg/L Fe3+).
    Fig. 9.  Selective inhibition behavior of Fe3+ on the galena surface.
    $$ \mathrm{P}\mathrm{b}\mathrm{S}+{8\mathrm{F}\mathrm{e}}^{3+}+4{\mathrm{H}}_{2}\mathrm{O}\to \mathrm{P}\mathrm{b}\mathrm{S}{\mathrm{O}}_{4}+8{\mathrm{F}\mathrm{e}}^{2+}+8{\mathrm{H}}^+ $$ (1)

    This study demonstrated the effect of Fe3+ on the flotation of chalcopyrite and galena as well as the galena inhibition mechanism of Fe3+ to present a new strategy for efficient chalcopyrite–galena separation. On the basis of the results and discussion presented above, the primary conclusions can be summarized as follows.

    (1) Fe3+ exerts a significant inhibitory effect on the floatability of galena while showing a negligible influence on the flotation of chalcopyrite when using IPETC as a collector.

    (2) The interaction between Fe3+ and galena is a redox reaction in which the PbS present on the surface of galena is oxidized into PbSO4, whereas Fe3+ is reduced into Fe2+.

    (3) The formation of compact and nanosized PbSO4 particles on the galena surface effectively prevents the adsorption of collector molecules, resulting in a hydrophilic galena surface.

    These findings provide a promising avenue for the efficient and clean separation of chalcopyrite and galena.

    Acknowledgements: This work was financially supported by the National Natural Science Foundation of China (Nos. 52204298 and 52004335), the National Key R&D Program of China (Nos. 2022YFC2904502 and 2022YFC2904501), the Major Science and Technology Projects in Yunnan Province (No. 202202AB080012), and the Science Research Initiation Fund of Central South University (No. 202044019).

    The authors declare that there is no conflict of interest regarding the publication of this paper.

  • [1]
    W. Guo, B. Feng, J.X. Peng, W.P. Zhang, and X.W. Zhu, Depressant behavior of tragacanth gum and its role in the flotation separation of chalcopyrite from talc, J. Mater. Res. Technol., 8(2019), No. 1, p. 697. doi: 10.1016/j.jmrt.2018.05.015
    [2]
    Q. Zhang, S.M. Wen, Q.C. Feng, and H. Wang, Enhanced sulfidization of azurite surfaces by ammonium phosphate and its effect on flotation, Int. J. Miner. Metall. Mater., 29(2022), No. 6, p. 1150. doi: 10.1007/s12613-021-2379-y
    [3]
    J.S. Yu, R.Q. Liu, L. Wang, W. Sun, H. Peng, and Y.H. Hu, Selective depression mechanism of ferric chromium lignin sulfonate for chalcopyrite–galena flotation separation, Int. J. Miner. Metall. Mater., 25(2018), No. 5, p. 489. doi: 10.1007/s12613-018-1595-6
    [4]
    X.M. Qiu, H.Y. Yang, G.B. Chen, S.P. Zhong, C.K. Cai, and B.B. Lan, Inhibited mechanism of carboxymethyl cellulose as a galena depressant in chalcopyrite and galena separation flotation, Miner. Eng., 150(2020), art. No. 106273. doi: 10.1016/j.mineng.2020.106273
    [5]
    A.A. Nikolaev and B.E. Goryachev, Thermodynamic and flotation analysis of influence exerted by chromate ions on separation of galena and chaclopyrite in alkaline media, J. Min. Sci., 43(2007), No. 6, p. 670. doi: 10.1007/s10913-007-0074-7
    [6]
    X. Chen, G.H. Gu, and Z.X. Chen, Seaweed glue as a novel polymer depressant for the selective separation of chalcopyrite and galena, Int. J. Miner. Metall. Mater., 26(2019), No. 12, p. 1495. doi: 10.1007/s12613-019-1848-z
    [7]
    Q. Liu and Y.H. Zhang, Effect of calcium ions and citric acid on the flotation separation of chalcopyrite from galena using dextrin, Miner. Eng., 13(2000), No. 13, p. 1405. doi: 10.1016/S0892-6875(00)00122-9
    [8]
    Z.J. Piao, D.Z. Wei, Z.L. Liu, W.G. Liu, S.L. Gao, and M.Y. Li, Selective depression of galena and chalcopyrite by O,O-bis(2,3-dihydroxypropyl) dithiophosphate, Trans. Nonferrous Met. Soc. China, 23(2013), No. 10, p. 3063. doi: 10.1016/S1003-6326(13)62834-4
    [9]
    J.M. Li, K.W. Song, D.W. Liu, et al., Hydrolyzation and adsorption behaviors of SPH and SCT used as combined depressants in the selective flotation of galena from sphalerite, J. Mol. Liq., 231(2017), p. 485. doi: 10.1016/j.molliq.2017.02.035
    [10]
    R.Z. Liu, W.Q. Qin, F. Jiao, et al., Flotation separation of chalcopyrite from galena by sodium humate and ammonium persulfate, Trans. Nonferrous Met. Soc. China, 26(2016), No. 1, p. 265. doi: 10.1016/S1003-6326(16)64113-4
    [11]
    P. Huang, L. Wang, and Q. Liu, Depressant function of high molecular weight polyacrylamide in the xanthate flotation of chalcopyrite and galena, Int. J. Miner. Process., 128(2014), p. 6. doi: 10.1016/j.minpro.2014.02.004
    [12]
    P. Huang, M.L. Cao, and Q. Liu, Selective depression of sphalerite by chitosan in differential PbZn flotation, Int. J. Miner. Process., 122(2013), p. 29. doi: 10.1016/j.minpro.2013.04.010
    [13]
    P. Huang, M.L. Cao, and Q. Liu, Adsorption of chitosan on chalcopyrite and galena from aqueous suspensions, Colloids Surf. A: Physicochem. Eng. Aspects, 409(2012), p. 167. doi: 10.1016/j.colsurfa.2012.06.016
    [14]
    A. López-Valdivieso, L.A. Lozano-Ledesma, A. Robledo-Cabrera, and O.A. Orozco-Navarro, Carboxymethylcellulose (CMC) as PbS depressant in the processing of Pb–Cu bulk concentrates. Adsorption and floatability studies, Miner. Eng., 112(2017), p. 77. doi: 10.1016/j.mineng.2017.07.012
    [15]
    Z.G. Yin, W. Sun, Y.H. Hu, et al., Synthesis of acetic acid-​[(hydrazinylthioxome​thyl)​thio]-sodium and its application on the flotation separation of molybdenite from galena, J. Ind. Eng. Chem., 52(2017), p. 82. doi: 10.1016/j.jiec.2017.03.027
    [16]
    X.R. Zhang, Z.B. Qian, G.B. Zheng, Y.G. Zhu, and W.G. Wu, The design of a macromolecular depressant for galena based on DFT studies and its application, Miner. Eng., 112(2017), p. 50. doi: 10.1016/j.mineng.2017.07.007
    [17]
    J.H. Chen, X.H. Long, L.H. Lan, and Q. He, Thermodynamics and density functional theory study of potassium dichromate interaction with galena, Int. J. Miner. Metall. Mater., 21(2014), No. 10, p. 947. doi: 10.1007/s12613-014-0994-6
    [18]
    R.P. Liao, S.M. Wen, Q.C. Feng, J.S. Deng, and H. Lai, Activation mechanism of ammonium oxalate with pyrite in the lime system and its response to flotation separation of pyrite from arsenopyrite, Int. J. Miner. Metall. Mater., 30(2023), No. 2, p. 271. doi: 10.1007/s12613-022-2505-5
    [19]
    G. Hong, J. Choi, Y. Han, et al., Relationship between surface characteristics and floatability in representative sulfide minerals: Role of surface oxidation, Mater. Trans., 58(2017), No. 7, p. 1069. doi: 10.2320/matertrans.M2017014
    [20]
    B. Feng, X.K. Jiao, H.H. Wang, J.X. Peng, and G. Yang, Improving the separation of chalcopyrite and galena by surface oxidation using hydroxyethyl cellulose as depressant, Miner. Eng., 160(2021), art. No. 106657. doi: 10.1016/j.mineng.2020.106657
    [21]
    D.W. Wang, F. Jiao, W.Q. Qin, and X.J. Wang, Effect of surface oxidation on the flotation separation of chalcopyrite and galena using sodium humate as depressant, Sep. Sci. Technol., 53(2018), No. 6, p. 961. doi: 10.1080/01496395.2017.1405042
    [22]
    S.A. Khoso, Y.H. Hu, F. Lü, Y. Gao, R.Q. Liu, and W. Sun, Xanthate interaction and flotation separation of H2O2-treated chalcopyrite and pyrite, Trans. Nonferrous Met. Soc. China, 29(2019), No. 12, p. 2604. doi: 10.1016/S1003-6326(19)65167-8
    [23]
    F. Jiang, L.M. Zhang, T. Yue, et al., Defect-boosted molybdenite-based co-catalytic Fenton reaction, Inorg. Chem. Front., 8(2021), No. 14, p. 3440. doi: 10.1039/D1QI00344E
    [24]
    A.P. Chandra, L. Puskar, D.J. Simpson, and A.R. Gerson, Copper and xanthate adsorption onto pyrite surfaces: Implications for mineral separation through flotation, Int. J. Miner. Process., 114-117(2012), p. 16. doi: 10.1016/j.minpro.2012.08.003
    [25]
    Y.L. Botero, A. Canales-Mahuzier, R. Serna-Guerrero, A. López-Valdivieso, M. Benzaazoua, and L.A. Cisternas, Physical-chemical study of IPETC and PAX collector’s adsorption on covellite surface, Appl. Surf. Sci., 602(2022), art. No. 154232. doi: 10.1016/j.apsusc.2022.154232
    [26]
    P. Forson, W. Skinner, and R. Asamoah, Decoupling pyrite and arsenopyrite in flotation using thionocarbamate collector, Powder Technol., 385(2021), p. 12. doi: 10.1016/j.powtec.2021.02.057
    [27]
    Y.H. Zhang, Z. Cao, Y.D. Cao, and C.Y. Sun, FTIR studies of xanthate adsorption on chalcopyrite, pentlandite and pyrite surfaces, J. Mol. Struct., 1048(2013), p. 434. doi: 10.1016/j.molstruc.2013.06.015
    [28]
    K. Laajalehto, J. Leppinen, I. Kartio, and T. Laiho, XPS and FTIR study of the influence of electrode potential on activation of pyrite by copper or lead, Colloids Surf. A: Physicochem. Eng. Aspects, 154(1999), No. 1-2, p. 193. doi: 10.1016/S0927-7757(98)00897-8
    [29]
    B.C. Wu, G.H. Gu, S. Deng, D.H. Liu, and X.X. Xiong, Efficient natural pyrrhotite activating persulfate for the degradation of O-isopropyl-N-ethyl thionocarbamate: Iron recycle mechanism and degradation pathway, Chemosphere, 224(2019), p. 120. doi: 10.1016/j.chemosphere.2019.02.062
    [30]
    H. Wang, J.C. Wang, Q.P. Zou, W.Q. Liu, C.Q. Wang, and W.Q. Huang, Surface treatment using potassium ferrate for separation of polycarbonate and polystyrene waste plastics by froth flotation, Appl. Surf. Sci., 448(2018), p. 219. doi: 10.1016/j.apsusc.2018.04.091
    [31]
    W.Q. Qin, X.J. Wang, L.Y. Ma, et al., Electrochemical characteristics and collectorless flotation behavior of galena: With and without the presence of pyrite, Miner. Eng., 74(2015), p. 99. doi: 10.1016/j.mineng.2015.01.010
    [32]
    G. De Giudici, A. Rossi, L. Fanfani, and P. Lattanzi, Mechanisms of galena dissolution in oxygen-saturated solutions: Evaluation of pH effect on apparent activation energies and mineral-water interface, Geochim. Cosmochim. Acta, 69(2005), No. 9, p. 2321. doi: 10.1016/j.gca.2004.12.003
    [33]
    Y.L. Mikhlin, A.A. Karacharov, and M.N. Likhatski, Effect of adsorption of butyl xanthate on galena, PbS, and HOPG surfaces as studied by atomic force microscopy and spectroscopy and XPS, Int. J. Miner. Process., 144(2015), p. 81. doi: 10.1016/j.minpro.2015.10.004
    [34]
    X. Ma, Y. Hu, H. Zhong, S. Wang, G.Y. Liu, and G. Zhao, A novel surfactant S-benzoyl-N,N-diethyldithiocarbamate synthesis and its flotation performance to galena, Appl. Surf. Sci., 365(2016), p. 342. doi: 10.1016/j.apsusc.2016.01.048
    [35]
    H.Y. Xie, Y.L. Jin, P. Zhang, et al., Surface modification mechanism of galena with H2SO4 and its effect on flotation separation performance, Appl. Surf. Sci., 579(2022), art. No. 152129. doi: 10.1016/j.apsusc.2021.152129
    [36]
    M.F. Liu, C.Y. Zhang, B. Hu, et al., Enhancing flotation separation of chalcopyrite and galena by the surface synergism between sodium sulfite and sodium lignosulfonate, Appl. Surf. Sci., 507(2020), art. No. 145042. doi: 10.1016/j.apsusc.2019.145042
    [37]
    G.K. Parker, R. Woods, and G.A. Hope, Raman investigation of chalcopyrite oxidation, Colloids Surf. A: Physicochem. Eng. Aspects, 318(2008), No. 1-3, p. 160. doi: 10.1016/j.colsurfa.2007.12.030
    [38]
    B. Li, L. Huang, M.Z. Zhong, Z.M. Wei, and J.B. Li, Electrical and magnetic properties of FeS2 and CuFeS2 nanoplates, RSC Adv., 5(2015), No. 111, p. 91103. doi: 10.1039/C5RA16918F
    [39]
    M.D. Lane, Mid-infrared emission spectroscopy of sulfate and sulfate-bearing minerals, Am. Mineral., 92(2007), No. 1, p. 1. doi: 10.2138/am.2007.2170
    [40]
    M. Monneron-Gyurits, E. Joussein, A. Courtin-Nomade, et al., A fast one-pot synthesize of crystalline anglesite by hydrothermal synthesis for environmental assessment on pure phase, Environ. Sci. Pollut. Res. Int., 29(2022), No. 12, p. 17373. doi: 10.1007/s11356-021-17011-6
    [41]
    B. Han, A.J. Xie, Q.B. Yu, F.Z. Huang, Y.H. Shen, and L. Zhu, Synthesis of PbSO4 crystals by hydrogel template on postprocessing strategy for secondary pollution, Appl. Surf. Sci., 261(2012), p. 623. doi: 10.1016/j.apsusc.2012.08.069
    [42]
    G. Falgayrac, S. Sobanska, J. Laureyns, and C. Brémard, Heterogeneous chemistry between PbSO4 and calcite microparticles using Raman microimaging, Spectrochim. Acta Part A: Mol. Biomol. Spectrosc., 64(2006), No. 5, p. 1095. doi: 10.1016/j.saa.2005.11.032
    [43]
    L. Yu, Q.J. Liu, S.M. Li, J.S. Deng, B. Luo, and H. Lai, Depression mechanism involving Fe3+ during arsenopyrite flotation, Sep. Purif. Technol., 222(2019), p. 109. doi: 10.1016/j.seppur.2019.04.007
    [44]
    P.X. Li, G. Zhang, W.J. Zhao, G. Han, and Q.C. Feng, Interaction mechanism of Fe3+ with smithsonite surfaces and its response to flotation performance, Sep. Purif. Technol., 291(2022), art. No. 121001. doi: 10.1016/j.seppur.2022.121001
  • Relative Articles

    [1]Feng Zhang, Chenyang Zhang, Linlin Wu, Wei Sun, Hongliang Zhang, Jianhua Chen, Yong Pei, Songjiang Li. Depression mechanism of sulfite ions on sphalerite and Pb2+ activated sphalerite in the flotation separation of galena from sphalerite[J]. International Journal of Minerals, Metallurgy and Materials, 2025, 32(2): 335-345. doi: 10.1007/s12613-024-2936-2
    [2]Haiyun Xie, Jialing Chen, Pei Zhang, Likun Gao, Dianwen Liu, Luzheng Chen. Separation of galena and chalcopyrite using the difference in their surface acid corrosion characteristics[J]. International Journal of Minerals, Metallurgy and Materials, 2023, 30(11): 2157-2168. doi: 10.1007/s12613-023-2654-1
    [3]Xiong Chen, Guo-hua Gu, Zhi-xiang Chen. Seaweed glue as a novel polymer depressant for the selective separation of chalcopyrite and galena[J]. International Journal of Minerals, Metallurgy and Materials, 2019, 26(12): 1495-1503. doi: 10.1007/s12613-019-1848-z
    [4]Zhi-li Li, Feng Rao, Shao-xian Song, Yan-mei Li, Wen-biao Liu. Slime coating of kaolinite on chalcopyrite in saline water flotation[J]. International Journal of Minerals, Metallurgy and Materials, 2018, 25(5): 481-488. doi: 10.1007/s12613-018-1594-7
    [5]Xiong Chen, Guo-hua Gu, Li-juan Li, Ren-feng Zhu. The selective effect of food-grade guar gum on chalcopyrite-monoclinic pyrrhotite separation using mixed aerofloat (CSU11) as collector[J]. International Journal of Minerals, Metallurgy and Materials, 2018, 25(10): 1123-1131. doi: 10.1007/s12613-018-1663-y
    [6]Jin-sheng Yu, Run-qing Liu, Li Wang, Wei Sun, Hong Peng, Yue-hua Hu. Selective depression mechanism of ferric chromium lignin sulfonate for chalcopyrite–galena flotation separation[J]. International Journal of Minerals, Metallurgy and Materials, 2018, 25(5): 489-497. doi: 10.1007/s12613-018-1595-6
    [7]Farhad Khorramshahi Mohammadabad, Sina Hejazi, Jalil Vahdati khaki, Abolfazl Babakhani. Mechanochemical leaching of chalcopyrite concentrate by sulfuric acid[J]. International Journal of Minerals, Metallurgy and Materials, 2016, 23(4): 380-388. doi: 10.1007/s12613-016-1247-7
    [8]Gustavo Urbano, Isabel Lázaro, Israel Rodríguez, Juan Luis Reyes, Roxana Larios, Roel Cruz. Electrochemical and spectroscopic study of interfacial interactions between chalcopyrite and typical flotation process reagents[J]. International Journal of Minerals, Metallurgy and Materials, 2016, 23(2): 127-136. doi: 10.1007/s12613-016-1219-y
    [9]Ritayan Chatterjee, Shamik Chaudhuri, Saikat Kumar Kuila, Dinabandhu Ghosh. Structural, microstructural, and thermal characterizations of a chalcopyrite concentrate from the Singhbhum shear zone, India[J]. International Journal of Minerals, Metallurgy and Materials, 2015, 22(3): 225-232. doi: 10.1007/s12613-015-1065-3
    [10]Yu-ting Liang, Shan Zhu, Jun Wang, Chen-bing Ai, Wen-qing Qin. Adsorption and leaching of chalcopyrite by Sulfolobus metallicus YN24 cultured in the distinct energy sources[J]. International Journal of Minerals, Metallurgy and Materials, 2015, 22(6): 549-552. doi: 10.1007/s12613-015-1106-y
    [11]Run-lan Yu, Jing Liu, Jian-xi Tan, Wei-min Zeng, Li-juan Shi, Guo-hua Gu, Wen-qing Qin, Guan-zhou Qiu. Effect of pH values on the extracellular polysaccharide secreted by Acidithiobacillus ferrooxidans during chalcopyrite bioleaching[J]. International Journal of Minerals, Metallurgy and Materials, 2014, 21(4): 311-316. doi: 10.1007/s12613-014-0910-0
    [12]Jian-Hua Chen, Xian-Hao Long, Li-Hong Lan, Qian He. Thermodynamics and density functional theory study of potassium dichromate interaction with galena[J]. International Journal of Minerals, Metallurgy and Materials, 2014, 21(10): 947-954. doi: 10.1007/s12613-014-0994-6
    [13]Bijan Taheri, Mahmoud Abdollahy, Sied Ziaedin Shafaei Tonkaboni, Soheyla Javadian, Mohammadreza Yarahmadi. Dual effects of sodium sulfide on the flotation behavior of chalcopyrite: I. Effect of pulp potential[J]. International Journal of Minerals, Metallurgy and Materials, 2014, 21(5): 415-422. doi: 10.1007/s12613-014-0924-7
    [14]Kubra Onol, Muhlis Nezihi Saridede. Investigation on microwave heating for direct leaching of chalcopyrite ores and concentrates[J]. International Journal of Minerals, Metallurgy and Materials, 2013, 20(3): 228-233. doi: 10.1007/s12613-013-0717-4
    [15]Ying-bo Dong, Hai Lin, Kai-bin Fu, Xiao-fang Xu, Shan-shan Zhou. Bioleaching of two different types of chalcopyrite by Acidithiobacillus ferrooxidans[J]. International Journal of Minerals, Metallurgy and Materials, 2013, 20(2): 119-124. doi: 10.1007/s12613-013-0702-y
    [16]To-hyon Pak, Ti-chang Sun, Jue Kou, Chol-Ryong Huang. Physical chemistry mechanisms of CDR system in sulphide mineral flotation[J]. International Journal of Minerals, Metallurgy and Materials, 2012, 19(3): 192-198. doi: 10.1007/s12613-012-0537-y
    [17]H. Naderi, M. Abdollahy, N. Mostoufi, M. J. Koleini, S. A. Shojaosadati, Z. Manafi. Kinetics of chemical leaching of chalcopyrite from low grade copper ore: behavior of different size fractions[J]. International Journal of Minerals, Metallurgy and Materials, 2011, 18(6): 638-645. doi: 10.1007/s12613-011-0489-7
    [18]Alafara A. Baba, Folahan A. Adekola. Comparative analysis of the dissolution kinetics of galena in binary solutions of HCl/FeCl3 and HCl/H2O2[J]. International Journal of Minerals, Metallurgy and Materials, 2011, 18(1): 9-17. doi: 10.1007/s12613-011-0393-1
    [19]Li-li Cheng, Ti-chang Sun, Xian-ping Luo, Dian-zuo Wang. Kinetics of the electrochemical process of galena electrodes in the diethyldithiocarbamate solution[J]. International Journal of Minerals, Metallurgy and Materials, 2010, 17(6): 669-674. doi: 10.1007/s12613-010-0372-y
    [20]Muqing Qiu, Shuiying Xiong, Weimin Zhang. Efficacy of chalcopyrite bioleaching using a pure and a mixed bacterium[J]. International Journal of Minerals, Metallurgy and Materials, 2006, 13(1): 7-10. doi: 10.1016/S1005-8850(06)60004-2
  • Periodical cited type(1)

    [1]Bingbing Li, Jieqi Chen, Dejian Cheng, et al. Small PbS Nanocubes Embedded in Carbon Shells for Na Storage. ACS Applied Nano Materials, 2024. https://doi.org/10.1021/acsanm.4c02993

    Other cited types(0)

  • 加载中

Catalog

    通讯作者: 陈斌, bchen63@163.com
    • 1. 

      沈阳化工大学材料科学与工程学院 沈阳 110142

    1. 本站搜索
    2. 百度学术搜索
    3. 万方数据库搜索
    4. CNKI搜索

    Figures(9)

    Share Article

    Article Metrics

    Article Views(1305) PDF Downloads(41) Cited by()
    Proportional views

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return