Gai-rong Wang, Hong-ying Yang, Yuan-yuan Liu, Lin-lin Tong, and Ali Auwalu, Study on the mechanical activation of malachite and the leaching of complex copper ore in the Luanshya mining area, Zambia, Int. J. Miner. Metall. Mater., 27(2020), No. 3, pp. 292-300. https://doi.org/10.1007/s12613-019-1856-z
Cite this article as:
Gai-rong Wang, Hong-ying Yang, Yuan-yuan Liu, Lin-lin Tong, and Ali Auwalu, Study on the mechanical activation of malachite and the leaching of complex copper ore in the Luanshya mining area, Zambia, Int. J. Miner. Metall. Mater., 27(2020), No. 3, pp. 292-300. https://doi.org/10.1007/s12613-019-1856-z
Research Article

Study on the mechanical activation of malachite and the leaching of complex copper ore in the Luanshya mining area, Zambia

+ Author Affiliations
  • Corresponding author:

    Hong-ying Yang    E-mail: yanghy@smm.neu.edu.cn

  • Received: 19 March 2019Revised: 15 May 2019Accepted: 3 June 2019Available online: 17 December 2019
  • Mechanical activation (MA) of malachite was carried out by dry planetary grinding (DPG) and wet Isa grinding (WIG) methods. When the rotational speed was increased to 400 r/min in DPG, the specific surface area of malachite reached the maximum and the particle size reached the minimum of 0.7–100 μm. Agglomeration occurred between mineral particles when the rotational speed was increased to 580 r/min in DPG. However, no agglomeration was observed among particles with sizes 0.4–3 μm in WIG. X-ray diffraction analysis showed that, at a 580 r/min rotational speed in DPG, the amorphization degree of malachite was 53.12%, whereas that in WIG was 71.40%, indicating that MA led to amorphization and distortion of crystal structures. In addition, in the Fourier transform infrared (FT-IR) spectra of activated malachite, the bands associated with –OH, ${\rm CO}_3^{2-} $and metal lattice vibrations of Cu–O and Cu–OH were weakened, and a new H–O–H bending mode and peaks of gaseous CO2 appeared, indicating that MA decreased the band energy, enhanced dihydroxylation, and increased the chemical reactivity of the malachite. Furthermore, the leaching behavior of copper ore was greatly improved by MA.
  • According to their proportion of oxide minerals and sulfide minerals, copper ores can be categorized as sulfide ores containing less than 10wt% copper oxide, mixed ores containing 10wt%–30wt% copper oxide, or oxide ores containing more than 30wt% copper oxide [1]. With the continuous consumption of copper sulfide ores, the development and utilization of copper oxide ores is becoming increasingly important. Malachite (Cu2(OH)2CO3) is one of the most important minerals in copper oxide ores. At present, complex copper oxide ores rich in malachite are used to produce copper through various hydrometallurgical processes, including acid leaching [24], ammonia leaching [57], and biological leaching [8]. Such hydrometallurgical processes can achieve substantial copper leaching rates but with the disadvantages of long leaching times, the generation of environmental pollution, and high consumption of reagents.

    Overcoming the aforementioned shortcomings associated with leaching malachite-rich complex copper oxide ores necessitates pretreatment before leaching. At present, the common pretreatment methods are mechanical activation (MA) and electromagnetic wave activation (microwaves, ultrasound, etc.) [9]. MA is a process that can increase the internal energy and reaction activity of minerals through mechanical forces such as friction, impact, collision, and shear, which lead to lattice distortion, lattice defects, and amorphization [1011]. MA leads to changes in certain properties of mineral particles, such as an increase in their specific surface area and structural disorder [12], an enhancement of strain [13], an increase in amorphization [10], alteration of their microtopography [14], and the formation of new phases more amenable to being leached [15]. Recent studies have shown that the process of MA generally includes two grinding modes: dry and wet grinding modes. The dry grinding mode mainly produces the milling effects of rubbing and crushing that cannot reduce small particles to finer particles [16]. However, the wet grinding mode relies on high-energy agitation to effectively drive small grinding media and form a large number of compressed rotating media layers that can create pressure and torsional forces. These two forces are much more effective for grinding small particles than the rubbing and crushing forces generated by the dry grinding mode [17]. For example, the X-ray diffraction (XRD) analysis of activated realgar revealed broadened diffraction peaks with diminished intensity after activation; in addition, the peak of an obviously amorphous phase appeared at 2θ = 16° [18]. In addition, Kleiv and Thornhill [19] reported that the specific surface area of olivine was substantially increased by MA, and that it was higher after combined activation by dry and wet modes (64.44 ± 2.13 m2/g) than after dry activation alone (26.28 ± 1.28 m2/g). Tkáčová and Stevulová [20] investigated the effects of activation of carbonates and quartz by dry and wet grinding in air and aqueous environments. Their results showed that the content of the X-ray-amorphous phase in the ground material increased with increasing grinding work although it remained unaffected by the grinding environment; in addition, the use of an aqueous environment resulted only in an enlarged specific surface area of the ground products. Dry grinding experiments involving quartz-based malachite mixtures were carried out for different grinding times by Tunç and Yildiz [21]. Their results showed obvious effects of the MA phenomenon, such as broadening of the XRD peaks, a reduction in XRD peak intensity, a shift of the XRD angles, and the generation of amorphous silicon. However, research on the characterization of activated malachite with regard to specific phases and crystal structures as well as the leaching effect have rarely been reported.

    In this work, the mechanism of MA and its effect on the physicochemical properties of malachite by dry planetary grinding (DPG) and wet Isa grinding (WIG) were investigated using XRD, scanning electron microscopy with energy-dispersive X-ray spectroscopy (SEM-EDS), and Fourier transform infrared (FT-IR) spectroscopy. The results of this study not only reveal a convenient method for obtaining malachite but also provide a theoretical basis and significant guidance for the leaching of malachite-rich complex copper oxide ores.

    The elements of malachite samples were analyzed by X-ray fluorescence (XRF) spectroscopy; the results are presented in Table 1. The main elements were Cu, Co, Fe, and Si. The quantitative analysis indicated that the sample contained 53.5wt% Cu with 5.3wt% C. Fig. 1 shows the XRD pattern, confirming the presence of malachite. In addition, the complex copper ores rich in malachite were obtained by open-pit mining from the Luanshya mining area, Zambia. The main Cu-bearing minerals were malachite, chrysocolla, and pseudo-malachite [22].

    Table 1.  XRF analysis results for the malachite sample wt%
    CuOCo2O3Fe2O3SiO2HfO2MgOAl2O3P2O5MnOOthers
    91.233.991.471.380.490.380.340.320.180.22
     | Show Table
    DownLoad: CSV
    Fig. 1.  XRD pattern of the malachite sample.

    In this study, the MA of malachite was performed in a dry planetary mill (QM3SP2 Planetary mill, China) and a wet Isa mill (ALC-1.5L mill, China). The dry planetary milling was conducted as follows: 15 g of malachite sample was combined with 180 g of zirconia grinding balls, where the mass ratio of grinding balls 5, 10, and 20 mm diameter was 3:4:5, respectively. The samples were mechanically activated at different rotation speeds from 200 to 580 r/min for 10 min (580 r/min was the rated speed). The conditions used for the wet Isa milling were as follows: 100 g of malachite sample and 3600 g of ceramic grinding balls (1 mm in diameter), and 800 mL of water were combined, and the samples were mechanically activated at 2000 r/min for 10 min.

    The Brunauer–Emmett–Teller (BET) specific surface area was determined via the low-temperature nitrogen adsorption method using a NOVA 1200e surface area and pore size analyzer. The values were calculated according to the BET theory. The micro-morphologies of the particles were characterized by SEM using an electron microscope (Shimadzu SSX-550) equipped with an EDS detector. The particle size distribution (PSD) was estimated using a laser particle sizer (Malvern Instruments Mastersizer 2000). The crystal structure analysis of malachite samples was conducted through XRD (SmartLab) analysis. The measurement conditions were as follows: Cu Kα1 radiation, curved height 0.42, scan range 10° ≤ 2θ ≤ 40°, step scan with a speed of 1°/min, U = 40 kV, I = 200 mA, T = 23.0°C. The chemical bonds of malachite samples were evaluated using FT-IR spectroscopy (VERTEX70) over the mid-infrared spectral range from 450 to 3500 cm−1.

    In this experiment, the leaching of nonactivated and activated copper ores rich in malachite were carried out in sulfuric acid. The leaching conditions were as follows: sulfuric acid 0.5 mol/L, temperature 30°C, solid-to-liquid ratio 50 g/L, stirring speed 300 r/min, and leaching time 120 min.

    The BET specific surface area (SBET) of nonactivated malachite samples and samples mechanically activated by different grinding methods are presented in Fig. 2. Compared with the BET specific surface area of the nonactivated malachite, that of activated malachite was greater, indicating that the particles were further refined. For the malachite activated by DPG, with increasing rotational speed, the BET specific surface area increased, reaching a maximum of 14.158 m2/g at a rotation speed of 400 r/min, which is 1.43 times greater than the BET surface area of the nonactivated sample. However, the BET specific surface area of malachite decreased unexpectedly when the rotational speed was increased to 580 r/min. The mineral samples adhered to the balls and sunk to the bottom at the highest grinding speed (580 r/min) in DPG, leading to agglomeration between the small mineral particles and to adsorption of the small particles onto the surface of the large particles [23]. In addition, a BET specific surface area of 25.280 m2/g was achieved in WIG, which is 2.56 times greater than that of the nonactivated sample. Although the rotational speed (2000 r/min) of the Isa mill was almost 3.45 times greater than that of the planetary mill (580 r/min), no agglomeration was observed between particles and the grains were much finer in the sample subjected to WIG. This result was attributed to the high-strength wet agitation of the Isa mill effectively dispersing the particles rather than inducing their agglomeration [24].

    Fig. 2.  BET specific surface area of malachite samples activated by different grinding methods: (a) nonactivated; (b) 200 r/min in DPG; (c) 400 r/min in DPG; (d) 580 r/min in DPG; (e) WIG.

    Fig. 3 shows the SEM-EDS images of malachite samples activated by different grinding methods. As shown in Fig. 3(a), the nonactivated malachite consisted of compact particles with different sizes and patterns. The patterns were irregular in shape, with a clear rhombus angle and smooth surface. Additionally, the uneven sizes of the malachite particles indicated that they lacked homogeneity. When the DPG rotational speed was increased to 200 r/min, the morphological features of the malachite samples were obviously changed. These changes were reflected by a sharp decrease in particle size and a roughened surface, as shown in Fig. 3(b). As the rotational speed was increased to 580 r/min in DPG, the malachite samples were substantially transformed from angular shapes to sphere-like structures. Nevertheless, serious agglomeration of the malachite particles is appeared at rotational speed of 580 r/min in Fig. 3(c), consistent with the results of the BET specific surface area shown in Fig. 2(d).

    Fig. 3.  SEM images and EDS analyses of malachite samples activated by different grinding methods: (a) nonactivated; (b) 200 r/min in DPG; (c) 580 r/min in DPG; (d) WIG. Spectra EDS1, EDS2, and EDS3 correspond to points 1, 2, and 3, respectively, in SEM micrographs.

    Compared with the particles of malachite samples activated by DPG, those of the samples activated by WIG were extremely small and dispersed uniformly, with no agglomeration (Fig. 3(d)). This phenomenon is attributed to the fact that the high-strength wet agitation in WIG paves no way for the particles to mutually agglomerate. In addition, the EDS spectrogram reveals that no phase transformation of malachite occurred during either DPG or WIG.

    The PSD curves of the malachite samples activated by different grinding methods are presented in Fig. 4. The PSD curve of the nonactivated malachite shows a single, broad peak corresponding to a size range from 0.5 to 300 μm. However, the PSD curves for the samples activated by DPG show broad distributions with multiple peaks. When the rotational speed was 200 r/min, the particle size range was 0.4–250 μm. As the rotational speed was increased to 400 r/min, the particle size range still was 0.4–250 μm while the distribution of large particle decreased. As the rotational speed was increased further to 580 r/min, the particle size range increased to 0.4–500 μm. The abnormal PSD of the sample activated by DPG at 580 r/min is attributed to the higher grinding speed causing the mineral samples adhere to the balls and sink to the bottom, leading to agglomeration among the small mineral particles and to adsorption of the small particles onto the surface of the large particles, which led to larger particles. This result is consistent with the analysis of the BET specific surface area and the aforementioned particle micromorphology.

    Fig. 4.  Particle size distributions of malachite samples activated by different grinding methods: (a) nonactivated; (b) 200 r/min in DPG; (c) 400 r/min in DPG; (d) 580 r/min in DPG; (e) WIG.

    The results in Fig. 4 also show that the peak in the PSD curve of the sample subjected to WIG was single and narrow, corresponding to small, 0.4–3 μm particles. This result differs substantially from the single peak in the curve of the nonactivated malachite sample, which contained large, 10–300 μm particles, indicating that the micro refinement effect of WIG was remarkable.

    Fig. 5 represents the XRD patterns of malachite samples activated by different grinding methods. In addition, the corresponding crystal lattice parameters are presented in Table 2, which also includes the average apparent crystallite size D and the average maximum strain ε. The results in Table 2 indicate that the parameters a, b, c, V, and D gradually decreased as the rotational speed was increased to 400 r/min in DPG. At this speed, the parameters a, b, c, V, and D reached their minimum values, in accordance with the maximum specific surface area as well as the smallest particle size, as previously discussed. However, when the rotational speed was increased to 580 r/min, the values of parameters a, b, c, V, and D did not continue their previous decreasing trend but instead increased because of agglomeration of the mineral particles. Unexpectedly, the average maximum strain ε gradually increased from 0.022% for the nonactivated sample to 0.034% for the sample activated at 580 r/min, indicating that greater surface energy was produced at high rotational speeds in DPG, leading to a continuous increase in the particle lattice distortion. Moreover, the smallest values of parameters a, c, V, and D and the highest average maximum strain ε of 0.045% were achieved in WIG, which indicates that the effects of reducing the particle crystallinity and increasing the amorphization in WIG were far more significant than those in DPG.

    Table 2.  Crystallographic data and microstructural parameters of malachite samples activated by different grinding methods
    Grinding methodsa / nmb / nmc / nmβ / (°)V / nm3D / nmε / %
    Nonactivated0.950811.196440.3244698.62510.3649275.80.022
    200 r·min–1 in DPG0.950091.196150.3244498.69330.3645075.50.023
    400 r·min–1 in DPG0.949501.195810.3244098.61810.3644874.00.032
    580 r·min–1 in DPG0.950231.195720.3243798.69460.3645874.50.034
    WIG0.949011.196030.3243598.66930.3643659.90.045
    Note: a, b, c, and β are unit-cell parameters; V is the unit-cell volume; D is the average apparent crystallite size; ε is the average maximum strain.
     | Show Table
    DownLoad: CSV
    Fig. 5.  XRD patterns of malachite samples activated by different grinding methods: (a) nonactivated; (b) 200 r/min in DPG; (c) 400 r/min in DPG; (d) 580 r/min in DPG; (e) WIG.

    The XRD patterns in Fig. 5 clearly show that the XRD diffraction peak characteristics changed, but no phase transition was observed for the malachite samples activated by the different grinding methods. With increasing rotational speed in DPG, the XRD diffraction peaks of the activated samples generally broadened and their intensities decreased. This phenomenon was more evident in WIG, indicating that the distortion of the crystal structure and increase in the amorphization degree in WIG were more obvious than those in DPG. Additionally, the amorphization degree of malachite was calculated by the equation [2526].

    Fig. 6 shows the amorphization degrees of malachite activated by different grinding methods. These results indicate that the amorphization degrees of malachite increased gradually with increasing rotational speed during DPG. When the rotational speed was at 580 r/min in DPG, the amorphization degrees was 53.12%, whereas it was 71.40% in WIG, further indicating that the effect of MA by WIG was better than that by DPG.

    Fig. 6.  Amorphization degrees of malachite samples activated by different grinding methods: (a) nonactivated; (b) 200 r/min in DPG; (c) 400 r/min in DPG; (d) 580 r/min in DPG; (e) WIG.

    Fig. 7 shows the FT-IR spectra of malachite samples activated by different grinding methods. The main absorption peaks in the FT-IR spectra are shown in Table 3 [27]. The spectra in Fig. 7 show that the intensive bands of malachite samples can be divided into four regions: 2800–3500 cm−1, 1000–1500 cm−1, 650–950 cm−1, and <650 cm−1, respectively. The centered bands of 3404.40, 3313.24, 1490.92, 1386.69, 1096.27, 1046.17, 819.73, 748.04, 712.03, 569.98, and 523.53 cm−1 were determined for a nonactivated malachite sample. However, some bands clearly changed in intensity, number, and position in the spectra of the activated samples, especially the sample activated by DPG at 580 r/min and that activated by WIG.

    Table 3.  Assignment of main absorption peaks in the infrared spectra of malachite samples
    Grinding methodsν(–OH)δ(–OH)ν1(${\rm CO}_3^{2-} $)ν2(${\rm CO}_3^{2-} $)ν3(${\rm CO}_3^{2-} $)ν4(${\rm CO}_3^{2-} $)Cu–OCu–OH
    Nonactivated3404.403313.241046.17870.151096.27819.731490.921386.69748.04712.03569.98523.52
    200 r·min–1 in DPG3404.493313.951046.85872.771096.38820.691491.351387.41748.36711.92571.59523.66
    400 r·min–1 in DPG3404.513314.681047.37874.941096.40820.921492.931388.36748.86712.09571.24524.13
    580 r·min–1 in DP3404.293315.771047.40876.421096.77821.011493.061389.24749.34712.40572.52524.41
    WIG3404.773317.311048.21878.481096.92821.171494.791389.58750.25712.86583.76524.89
    Ref. [27]340033201045875109582015001400748710570, 580522
     | Show Table
    DownLoad: CSV
    Fig. 7.  FT-IR spectra of malachite samples activated by different grinding methods: (a) nonactivated; (b) 200 r/min in DPG; (c) 400 r/min in DPG; (d) 580 r/min in DPG; (e) WIG.

    Fig. 7 shows that, as the rotational speed was increased to 400 r/min in DPG, the intensity and area of most of the absorption bands remained unchanged; however, the bands changed when the rotational speed was increased to 580 r/min. This observation was attributed to the activity of the mineral group being enhanced by the high-energy grinding, resulting in instability of the chemical bonds [28]. In addition, the intensity and area of the absorption bands increased substantially in the spectra of the sample activated by WIG compared with those in the spectra of the samples activated by DPG and the nonactivated samples. That is, the intensities of these bands were enhanced with increasing mineral disorder induced by grinding [29].

    Goldsmith and Ross [27] have reported that the symmetrical vibration bands ν(–OH) groups appear at 3400 and 3320 cm−1 [27]. Tunç and Yildiz [21] mechanically activated impure malachite containing quartz and showed that the two –OH bands disappeared after activation. However, because of the high purity of malachite samples used in the present study, the two bands were still present at 3404.40 and 3313.24 cm−1 in the spectra of the activated malachite. We also found that the position of the –OH symmetrical vibration band remained almost unchanged in the spectra of DPG and WIG compared with that in the spectrum of the nonactivated malachite. Nevertheless, the positions of the two deformation vibration bands δ(–OH) at 1046.17 and 870.15 cm−1 slightly shifted to 1047.40 and 876.42 cm−1 in the spectrum of the sample activated by DPG at 580 r/min and to 1048.21 and 878.48 cm−1 in the spectrum of the sample activated by WIG, respectively. The center-shift of a band to higher frequency indicates a decrease in bond strength [30], suggesting that the –OH bond was weakened and that the free –OH content was increased. In the meantime, a new band attributed to the H–O–H bending mode gradually emerged at 1780.78 cm−1, whereas no such band was observed in the spectrum of the nonactivated malachite [31], confirming that the structure of –OH was converted to free water via DPG and WIG [32]. Dehydroxylation is mediated by the interaction of two hydroxyl groups in a two-step proton transfer process that results in the formation of a water molecule and in a chemically bonded oxygen atom being left in the form of a superoxide anion in the lattice. This reaction process is represented chemically in Eqs. (1) and (2) [33]:

    $$ {\rm{O}}{{\rm{H}}^ - } \rightleftharpoons {{\rm{H}}^ + } + {{\rm{O}}^{2 - }} $$ (1)
    $$ {{\rm{H}}^ + } + {\rm{O}}{{\rm{H}}^ - } \rightleftharpoons {{\rm{H}}_2}{\rm{O}} $$ (2)

    The position of the symmetric stretching vibration ν1(${\rm CO}_3^{2-} $) at 1096.27 cm−1 was unchanged in the spectra of the samples activated by WIG and by DPG at 580 r/min. However, the unsymmetrical stretching vibrations ν3(${\rm CO}_3^{2-} $) shifted from 1386.69 and 1490.92 cm−1 in the spectra of nonactivated malachite to 1389.24 and 1493.06 cm−1 in the spectrum of the sample activated by DPG at 580 r/min and to 1389.58 and 1494.79 cm−1 in the spectrum of the sample activated by WIG [34]. In addition, the peak positions of the in-plane bending vibration ν2(${\rm CO}_3^{2-} $) at 819.73 cm−1 and the out-of-plane bending vibrations ν4(${\rm CO}_3^{2-} $) at 748.04 and 712.03 cm−1 shifted to 821.17 and to 750.25 and 712.86 cm−1 in the spectrum of the sample activated by WIG, respectively. The shift of a band center to higher frequencies indicates a decrease in bond strength [30]. Therefore, in combination with the results showing that the bands of ${\rm CO}_3^{2-} $ shifted to a higher position and the intensity increased, these results indicate that the absorption bands associated with ${\rm CO}_3^{2-} $ groups became gradually weaker. Moreover, we observed that a new band at 2366.71 cm−1 appeared in the spectra of the samples activated by DPG at 200, 400, and 580 r/min. The new band was attributed to gaseous CO2, which was generated by decomposition of the atmosphere or ${\rm CO}_3^{2-} $ groups [35], as shown in Eq. (3). However, the new band of gaseous CO2 nearly disappeared in the spectrum of the sample activated by WIG, leading to the presumption that the gaseous CO2 was dissolved in water, as shown in Eq. (4):

    $$ {\rm{C}}{{\rm{u}}_2}{\left( {{\rm{OH}}} \right)_2}{\rm{C}}{{\rm{O}}_3} = 2{\rm{CuO}} + {{\rm{H}}_2}{\rm{O}} + {\rm{C}}{{\rm{O}}_2} \uparrow $$ (3)
    $$ {\rm{C}}{{\rm{O}}_2} + {{\rm{H}}_2}{\rm{O}} = {{\rm{H}}_2}{\rm{C}}{{\rm{O}}_3} $$ (4)

    The absorption bands of metal lattice vibrations of Cu–O and Cu–OH appear at 569.98 and 523.52 cm−1, respectively. Their positions shifted to 572.52 and 524.41 cm−1 in the spectrum of the sample activated by DPG at 580 r/min and to 583.76 and 524.89 cm−1 in the spectrum of the sample activated by WIG, thereby demonstrating that the effect of MA was more remarkable in WIG than in DPG.

    Fig. 8 shows the leaching results for malachite-rich complex copper ore activated by different grinding methods and leached in sulfuric acid. The copper leaching rate obviously increased rapidly before 40 min and then increased slowly at 200 r/min in DPG, similar to the nonactivated ore. For the ores activated at 400 and 580 r/min, the leaching rate of copper increased rapidly before 10 min and then increased slowly. However, we found that the leaching rate of copper was extremely high before 2 min and then increased slowly in WIG. Although agglomerated particles were observed in the aforementioned analysis of the samples activated by DPG at 580 r/min, the leaching rate of ore activated at 580 r/min was slightly greater than that of the ore activated at 400 r/min, which is a consequence of a more distorted crystal structure and a greater amorphization degree of samples ground at higher speeds. Therefore, the order of copper leaching rate with respect to different grinding methods was achieved as follows: WIG > DPG > nonactivated, which well matches the results of the analyses of the specific surface area, PSD, and amorphization degree of the activated malachite.

    Fig. 8.  Leaching results of malachite-rich copper oxide ore activated by different grinding methods and leached in sulfuric acid: (a) nonactivated; (b) 200 r/min in DPG; (c) 400 r/min in DPG; (d) 580 r/min in DPG; (e) WIG.

    (1) The malachite samples were mechanically activated in DPG and WIG. The samples activated by WIG exhibited remarkable improvements in specific surface area, particle micromorphology, PSD, crystal structure, and chemical bonding.

    (2) Agglomeration occurred between the mineral particles when the rotational speed was increased to 580 r/min in DPG. However, no particle agglomeration was observed for the samples activated by WIG, which exhibited a particle size of 0.4–3 μm.

    (3) XRD analysis indicated that no phase transformation of the activated malachite occurred in either DPG or WIG; however, the amorphization of the samples gradually changed. When the rotational speed was at 580 r/min in DPG, the amorphization degree of malachite was 53.12%, whereas it was 71.40% in WIG, showing that MA leads to amorphization and to distortion of the malachite crystal structure.

    (4) In the FT-IR analysis, the absorption bands associated with –OH vibrations, ${\rm CO}_3^{2-} $ vibrations, and metal lattice vibrations of Cu–O and Cu–OH weakened and a new H–O–H bending mode as well as gaseous CO2 appeared on the activated malachite minerals, indicating that MA can decrease the band energy, enhance dihydroxylation, and increase chemical reactivity, especially when conducted by WIG.

    (5) The leaching properties of malachite-rich copper ore with respect to different grinding methods were greatly improved, especially the WIG, which well matches the results of the analyses of the specific surface area, PSD, and amorphization degree of the activated malachite.

    This work was financially supported by the Special Funds for the National Natural Science Foundation of China (No. U1608254), and the National Key R&D Program of China (No. 2018YFC1902002).

  • [1]
    Y.Q. Zhao, Treatment of Copper Oxide Ore, The Metallurgical Industry Press, Beijing, 1982.
    [2]
    O.N. Ata, S. Çolak, Z. Ekinci, and M. Çopur, Determination of the optimum conditions for leaching of malachite ore in H2SO4 solutions, Chem. Eng. Technol., 24(2001), No. 4, p. 409. doi: 10.1002/1521-4125(200104)24:4<409::AID-CEAT409>3.0.CO;2-0
    [3]
    J.S. Deng, S.M. Wen, Q. Yin, D.D. Wu, and Q.W. Sun, Leaching of malachite using 5-sulfosalicylic acid, J. Taiwan Inst. Chem. Eng., 71(2017), p. 20. doi: 10.1016/j.jtice.2016.11.013
    [4]
    D. Bingöl and M. Canbazoğlu, Dissolution kinetics of malachite in sulphuric acid, Hydrometallurgy, 72(2004), No. 1-2, p. 159. doi: 10.1016/j.hydromet.2003.10.002
    [5]
    D. Bingöl, M. Canbazoğlu, and S. Aydoğan, Dissolution kinetics of malachite in ammonia/ammonium carbonate leaching, Hydrometallurgy, 76(2005), No. 1-2, p. 55. doi: 10.1016/j.hydromet.2004.09.006
    [6]
    A. Künkül, M.M. Kocakerim, S. Yapici, and A. Demirbaǧ, Leaching kinetics of malachite in ammonia solutions, Int. J. Miner. Process., 41(1994), No. 3-4, p. 167. doi: 10.1016/0301-7516(94)90026-4
    [7]
    X. Wang, Q.Y. Chen, H.P. Hu, Z.L. Yin, and Z.L. Xiao, Solubility prediction of malachite in aqueous ammoniacal ammonium chloride solutions at 25°C, Hydrometallurgy, 99(2009), No. 3-4, p. 231. doi: 10.1016/j.hydromet.2009.08.011
    [8]
    K.B. Fu, H. Lin, Y.B. Dong, X.L. Mo, and H. Wang, The rules of bacteria in malachite leaching of low acidity, Met. Mine, 40(2011), No. 3, p. 69.
    [9]
    R.J. Ma, New development of hydrometallurgy, Hydrometall. China, 26(2007), No. 1, p. 1.
    [10]
    K. Tkáčová, P. Baláž, B. Mišura, V.E. Vigdergauz, and V.A. Chanturiya, Selective leaching of zinc from mechanically activated complex Cu Pb Zn concentrate, Hydrometallurgy, 33(1993), No. 3, p. 291. doi: 10.1016/0304-386X(93)90068-O
    [11]
    C. Zhang, Study on Mechanochemical Calorimeter and Mechanochemical Energy [Dissertation], Central South University, Changsha, 2008.
    [12]
    P. Baláž, Influence of solid state properties on ferric chloride leaching of mechanically activated galena, Hydrometallurgy, 40(1996), No. 3, p. 359. doi: 10.1016/0304-386X(95)00011-5
    [13]
    P. Baláž, Extractive Metallurgy of Activated Minerals, Elsevier, Amsterdam, 2000.
    [14]
    D. Tromans and J.A. Meech, Enhanced dissolution of minerals: microtopography and mechanical activation, Miner. Eng., 12(1999), No. 6, p. 609. doi: 10.1016/S0892-6875(99)00047-3
    [15]
    N.J. Welham, Enhanced dissolution of tantalite/columbite following milling, Int. J. Miner. Process., 61(2001), No. 3, p. 145. doi: 10.1016/S0301-7516(00)00032-6
    [16]
    P. Baláž, Mechanochemistry in Nanoscience and Minerals Engineering, Springer, Berlin, Heidelberg, 2008.
    [17]
    M.W. Gao, R.J. Holmes, and J. Pease, The latest developments in fine and ultrafine grinding technologies, [in] 23rd International Mineral Processing Congress, Istanbul, 2006, p. 30.
    [18]
    P. Baláž, M. Baláž, O. Shpotyuk, P. Demchenko, M. Vlček, M. Shopska, J. Briančin, Z. Bujňáková, Y. Shpotyuk, B. Selepová, and L. Balážová, Properties of arsenic sulphide (β-As4S4) modified by mechanical activation, J. Mater. Sci., 52(2017), No. 3, p. 1747. doi: 10.1007/s10853-016-0466-7
    [19]
    R.A. Kleiv and M. Thornhill, The effect of mechanical activation in the production of olivine surface area, Miner. Eng., 89(2016), p. 19. doi: 10.1016/j.mineng.2016.01.003
    [20]
    K. Tkáčová and N. Stevulová, Change in structure and enthalpy of carbonates and quartz accompanying grinding in air and aqueous environments, Powder Technol., 52(1987), No. 2, p. 161. doi: 10.1016/0032-5910(87)80146-8
    [21]
    T. Tunç and K. Yildiz, Structural alterations in mechanically activated malachite, Acta Phys. Pol. A, 125(2014), No. 2, p. 177. doi: 10.12693/APhysPolA.125.177
    [22]
    G.R. Wang, H.Y. Yang, L.L. Tong, and Y.Y. Liu, Research on process mineralogy of oxidized copper ore in Luanshya, Zambia, J. Northeastern Univ. Nat. Sci., 40(2019), No. 3, p. 350.
    [23]
    R. Mejdoub, H. Hammi, M. Khitouni, J.J. Suñol, and A. M'nif, The effect of prolonged mechanical activation duration on the reactivity of portland cement: Effect of particle size and crystallinity changes, Constr. Build. Mater., 152(2017), p. 1041. doi: 10.1016/j.conbuildmat.2017.07.008
    [24]
    S. Romeis, J. Schmidt, and W. Peukert, Mechanochemical aspects in wet stirred media milling, Int. J. Miner. Process., 156(2016), p. 24. doi: 10.1016/j.minpro.2016.05.018
    [25]
    P. Pourghahramani, E. Altin, M.R. Mallembakam, W. Peukert, and E. Forssberg, Microstructural characterization of hematite during wet and dry millings using rietveld and xrd line profile analyses, Powder Technol., 186(2008), No. 1, p. 9. doi: 10.1016/j.powtec.2007.10.027
    [26]
    S.M. Ohlberg and D.W. Strickler, Determination of percent crystallinity of partly devitrified glass by X-ray diffraction, J. Am. Ceram. Soc., 45(1962), No. 4, p. 170. doi: 10.1111/j.1151-2916.1962.tb11114.x
    [27]
    J.A. Goldsmith and S.D. Ross, The infra-red spectra of azurite and malachite, Spectrochim. Acta Part A, 24(1968), No. 12, p. 2131. doi: 10.1016/0584-8539(68)80273-9
    [28]
    M. Descamps and J.F. Willart, Perspectives on the amorphization/milling relationship in pharmaceutical materials, Adv. Drug Delivery Rev., 100(2016), p. 51. doi: 10.1016/j.addr.2016.01.011
    [29]
    P. Baláž, E. Turianicová, M. Fabián, R.A. Kleiv, J. Briančin, and A. Obut, Structural changes in olivine (Mg. Fe)2SiO4 mechanically activated in high-energy mills, Int. J. Miner. Process., 88(2008), No. 1-2, p. 1. doi: 10.1016/j.minpro.2008.04.001
    [30]
    T. Tunç, F. Apaydin, and K. Yildiz, Effects of mechanical activation on the structure of nickeliferous laterite, Acta Phys. Pol. A, 123(2013), No. 2, p. 349. doi: 10.12693/APhysPolA.123.349
    [31]
    Y.J. Wang, S.L. Pan, X.L. Hou, G. Liu, J.D. Wang, and D.Z. Jia, Non-centrosymmetric sodium borate: Crystal growth, characterization and properties on Na2B4O12H10, Solid State Sci., 12(2010), No. 10, p. 1726. doi: 10.1016/j.solidstatesciences.2010.07.021
    [32]
    Z. He, J. Zhou, Z.P. Lai, L.H. Yang, J.M. Liang, H. Long, and X.J. Ou, Quartz OSL dating of sand dunes of Late Pleistocene in the Mu Us Desert in northern China, Quat. Geochronol., 5(2010), No. 2-3, p. 102. doi: 10.1016/j.quageo.2009.02.011
    [33]
    E. Horváth, R.L. Frost, É. Makó, J. Kristóf, and T. Cseh, Thermal treatment of mechanochemically activated kaolinite, Thermochim. Acta, 404(2003), No. 1-2, p. 227. doi: 10.1016/S0040-6031(03)00184-9
    [34]
    R.L. Frost, W.N. Martens, L. Rintoul, E. Mahmutagic, and J.T. Kloprogge, Raman spectroscopic study of azurite and malachite at 298 and 77 K, J. Raman Spectrosc., 33(2002), No. 4, p. 252. doi: 10.1002/jrs.848
    [35]
    Y.Z. Xu, T. Jiang, M. Zhou, J. Wen, W.Y. Chen, and X.X. Xue, Effects of mechanical activation on physicochemical properties and alkaline leaching of boron concentrate, Hydrometallurgy, 173(2017), p. 32. doi: 10.1016/j.hydromet.2017.05.014
  • Relative Articles

    [1]Qicheng Feng, Wanming Lu, Han Wang, Qian Zhang. Mechanistic insights into stepwise activation of malachite for enhancing surface reactivity and flotation performance[J]. International Journal of Minerals, Metallurgy and Materials, 2024, 31(10): 2159-2172. doi: 10.1007/s12613-023-2793-4
    [2]Zhihao Shen, Shuming Wen, Han Wang, Yongchao Miao, Xiao Wang, Shengbing Meng, Qicheng Feng. Effect of dissolved components of malachite and calcite on surface properties and flotation behavior[J]. International Journal of Minerals, Metallurgy and Materials, 2023, 30(7): 1297-1309. doi: 10.1007/s12613-023-2606-9
    [3]Ping-chao Ke, Zhi-hong Liu, Lin Li. Synthesis, characterization, and property test of crystalline polyferric sulfate adsorbent used in treatment of contaminated water with a high As(III) content[J]. International Journal of Minerals, Metallurgy and Materials, 2018, 25(10): 1217-1225. doi: 10.1007/s12613-018-1674-8
    [4]Elham Hosseini, Fereshteh Rashchi, Abolghasem Ataie. Ti leaching from activated ilmenite-Fe mixture at different milling energy levels[J]. International Journal of Minerals, Metallurgy and Materials, 2018, 25(11): 1263-1274. doi: 10.1007/s12613-018-1679-3
    [5]Nathália Vieceli, Carlos A. Nogueira, Manuel F. C. Pereira, Fernando O. Durão, Carlos Guimarães, Fernanda Margarido. Optimization of an innovative approach involving mechanical activation and acid digestion for the extraction of lithium from lepidolite[J]. International Journal of Minerals, Metallurgy and Materials, 2018, 25(1): 11-19. doi: 10.1007/s12613-018-1541-7
    [6]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
    [7]Jiu-shuai Deng, Shu-ming Wen, Jian-ying Deng, Dan-dan Wu. Extracting copper from copper oxide ore by a zwitterionic reagent and dissolution kinetics[J]. International Journal of Minerals, Metallurgy and Materials, 2015, 22(3): 241-248. doi: 10.1007/s12613-015-1067-1
    [8]Jia-jie Li, Michael Hitch. Ultra-fine grinding and mechanical activation of mine waste rock using a high-speed stirred mill for mineral carbonation[J]. International Journal of Minerals, Metallurgy and Materials, 2015, 22(10): 1005-1016. doi: 10.1007/s12613-015-1162-3
    [9]Xiao-qing Niu, Ji-chun Chen. Morphology, crystal structure and hydration of calcined and modified anhydrite[J]. International Journal of Minerals, Metallurgy and Materials, 2014, 21(10): 1028-1032. doi: 10.1007/s12613-014-1004-8
    [10]Kolsoom Ahmadisoltansaraei, Javad Moghaddam. Preparation of NiO nanoparticles from Ni(OH)2·NiCO3·4H2O precursor by mechanical activation[J]. International Journal of Minerals, Metallurgy and Materials, 2014, 21(7): 726-735. doi: 10.1007/s12613-014-0964-z
    [11]Cui-hua Zhao, Bo-zeng Wu, Jian-hua Chen, Yu-qiong Li, Ye Chen. Electronic structure and properties of FeS2 with the space groups of Pa3 and P1[J]. International Journal of Minerals, Metallurgy and Materials, 2013, 20(7): 671-677. doi: 10.1007/s12613-013-0782-8
    [12]Lei Zhao, Bo-ping Zhang, Min Cui, Yan Li. Effect of KNbO3 content on the crystal structure and electrical properties of (1 − x)PbZr0.54Ti0.46O3-xKNbO3 piezoelectric ceramics[J]. International Journal of Minerals, Metallurgy and Materials, 2013, 20(3): 279-283. doi: 10.1007/s12613-013-0724-5
    [13]Li-hua Zhao, Li Zhan, Hao Bai, Da-qiang Cang. Effect of electric pulse on the crystal structure of solidified copper ingots[J]. International Journal of Minerals, Metallurgy and Materials, 2013, 20(6): 556-562. doi: 10.1007/s12613-013-0765-9
    [14]M. A. Shabani, M. Irannajad, A. R. Azadmehr. Investigation on leaching of malachite by citric acid[J]. International Journal of Minerals, Metallurgy and Materials, 2012, 19(9): 782-786. doi: 10.1007/s12613-012-0628-9
    [15]E. Darezereshki, M. Schaffie, M. Lotfalian, S. A. Seiedbaghery, M. Ranjbar. Use of mesophilic and thermophilic bacteria for the improvement of copper extraction from a low-grade ore[J]. International Journal of Minerals, Metallurgy and Materials, 2011, 18(2): 138-143. doi: 10.1007/s12613-011-0413-1
    [16]Zhong-lai Yi, Heng-hu Sun, Xiu-quan Wei, Chao Li. Iron ore tailings used for the preparation of cementitious material by compound thermal activation[J]. International Journal of Minerals, Metallurgy and Materials, 2009, 16(3): 355-358. doi: 10.1016/S1674-4799(09)60064-9
    [17]Pei Zhao, Boping Zhang, Ke Wang, Limin Zhang, Hailong Zhang. Effect of Li content on the microstructure and properties of lead-free piezoelectric(K0.5Na0.5)1-xLixNbO3 ceramics prepared by SPS[J]. International Journal of Minerals, Metallurgy and Materials, 2008, 15(3): 314-319. doi: 10.1016/S1005-8850(08)60059-6
    [18]Yuling Xie, Jiuhua Xu, Guangming Li, Zhiming Yang, Longsheng Yi. Characteristics and evolution of ore-forming fluids of the Chongjiang copper deposit in the Gangdise porphyry copper belt, Tibet[J]. International Journal of Minerals, Metallurgy and Materials, 2007, 14(2): 97-102. doi: 10.1016/S1005-8850(07)60020-6
    [19]Chengchang Jia, Xiuhu Guan, Xuekuan Su, Changheng Hou, Shimin Fan, Jun Zhao, Zizhang Xie. Microstructure and Properties of W-Cu Alloys Prepared with Mechanically Activated Powder[J]. International Journal of Minerals, Metallurgy and Materials, 2001, 8(2): 129-132.
    [20]Chengchang jia, Xiaosen Lü, Jun Zhao, Zizhang Xie. Reaction Sintering of Fe-Al Powder Mechanically Activated[J]. International Journal of Minerals, Metallurgy and Materials, 1998, 5(4): 233-236.
  • Periodical cited type(12)

    [1]Jayson Sime Jeremias, Maria Lourdes Dalida, Ming-Chun Lu. Copper recovery from treatment of simulated copper-iron wastewater using fluidized-bed homogeneous crystallization technology. Journal of Cleaner Production, 2024. https://doi.org/10.1016/j.jclepro.2024.140604
    [2]Zhong-lin DONG, Qian LI, Xue-fei RAO, et al. Economical and efficient recovery of Pd and Al2O3 from spent Pd/Al2O3 catalyst with integrated sodium roasting−reductive leaching process. Transactions of Nonferrous Metals Society of China, 2023, 33(7): 2245. https://doi.org/10.1016/S1003-6326(23)66256-9
    [3]Rafael Baieta, Vojtěch Ettler, Aleš Vaněk, et al. Smelter-derived soil contamination in Luanshya, Zambia. Science of The Total Environment, 2023, 867: 161405. https://doi.org/10.1016/j.scitotenv.2023.161405
    [4]Galina Kholodnaya, Denis Ponomarev, Roman Sazonov, et al. Effect of the Addition of a Buffer Gas during Pulsed Plasma-Chemical Synthesis on the Structural and Morphological Characteristics of CuxOy@SiO2 and CuxOy@TiO2 Nanocomposites. Journal of Materials Engineering and Performance, 2022, 31(10): 8027. https://doi.org/10.1007/s11665-022-06870-w
    [5]Baoqiang Yu, Jue Kou, Chunbao Sun, et al. Extraction of copper from copper-bearing biotite by ultrasonic-assisted leaching. International Journal of Minerals, Metallurgy and Materials, 2022, 29(2): 212. https://doi.org/10.1007/s12613-020-2132-y
    [6]Xiaofen Li, Wei Li, Yuntao Gao, et al. Effect of Mechanical Activation on the Leaching Process of Rare Earth Metal Yttrium in Deep Eutectic Solvents. Applied Sciences, 2022, 12(16): 8253. https://doi.org/10.3390/app12168253
    [7]Yunfeng Shi, Qi Zuo, Dan Liu, et al. Dissolution kinetics of malachite in trichloroacetic acid solution. Journal of the Iranian Chemical Society, 2022, 19(6): 2581. https://doi.org/10.1007/s13738-021-02479-3
    [8]Claudia Invernizzi, Lavinia de Ferri, Valeria Comite, et al. Correlation between surface roughness and spectral features in IR-reflection spectroscopy. Microchemical Journal, 2022, 172: 106874. https://doi.org/10.1016/j.microc.2021.106874
    [9]Qian Zhang, Shuming Wen, Qicheng Feng, et al. Enhanced sulfidization of azurite surfaces by ammonium phosphate and its effect on flotation. International Journal of Minerals, Metallurgy and Materials, 2022, 29(6): 1150. https://doi.org/10.1007/s12613-021-2379-y
    [10]Sheng-hua Yin, Lei-ming Wang, Xun Chen, et al. Agglomeration and leaching behaviors of copper oxides with different chemical binders. International Journal of Minerals, Metallurgy and Materials, 2021, 28(7): 1127. https://doi.org/10.1007/s12613-020-2081-5
    [11]Oluwasegun Samuel Odebiyi, Hao Du, Kayode Hassan Lasisi, et al. Effect of Ball Mill Parameters’ Variation on the Particles of a Mechanical Activation-Assisted Leaching: A Hydrometallurgical Mechanics. Materials Circular Economy, 2021, 3(1) https://doi.org/10.1007/s42824-021-00030-6
    [12]Kirill Karimov, Andrei Shoppert, Denis Rogozhnikov, et al. Effect of Preliminary Alkali Desilication on Ammonia Pressure Leaching of Low-Grade Copper–Silver Concentrate. Metals, 2020, 10(6): 812. https://doi.org/10.3390/met10060812

    Other cited types(0)

  • Created with Highcharts 5.0.7Amount of accessChart context menuView Count, PDF Downloads StatisticsView CountPDF Downloads2024-022024-032024-042024-052024-062024-072024-082024-092024-102024-112024-122025-01010203040Highcharts.com
    Created with Highcharts 5.0.7Chart context menuAccess Class DistributionView Count: 94.3 %View Count: 94.3 %PDF: 5.7 %PDF: 5.7 %View CountPDFHighcharts.com
    Created with Highcharts 5.0.7Chart context menuAccess Area Distribution其他: 1.6 %其他: 1.6 %Bangladesh: 0.1 %Bangladesh: 0.1 %Belgium: 0.5 %Belgium: 0.5 %Bulgaria: 0.4 %Bulgaria: 0.4 %Canada: 0.9 %Canada: 0.9 %Chile: 0.8 %Chile: 0.8 %China: 43.1 %China: 43.1 %Czech Republic: 0.5 %Czech Republic: 0.5 %Egypt: 0.5 %Egypt: 0.5 %France: 0.2 %France: 0.2 %Germany: 0.2 %Germany: 0.2 %Greece: 3.5 %Greece: 3.5 %Hong Kong, China: 1.0 %Hong Kong, China: 1.0 %Hungary: 0.5 %Hungary: 0.5 %India: 1.9 %India: 1.9 %Iran (ISLAMIC Republic Of): 1.2 %Iran (ISLAMIC Republic Of): 1.2 %Iraq: 0.2 %Iraq: 0.2 %Italy: 0.2 %Italy: 0.2 %Japan: 0.5 %Japan: 0.5 %Kazakhstan: 0.1 %Kazakhstan: 0.1 %Korea Republic of: 6.0 %Korea Republic of: 6.0 %Malaysia: 0.1 %Malaysia: 0.1 %Mexico: 0.2 %Mexico: 0.2 %Mongolia: 0.9 %Mongolia: 0.9 %Morocco: 0.4 %Morocco: 0.4 %Netherlands: 0.3 %Netherlands: 0.3 %Nigeria: 1.8 %Nigeria: 1.8 %Peru: 0.4 %Peru: 0.4 %Reserved: 0.8 %Reserved: 0.8 %Romania: 0.3 %Romania: 0.3 %Russian Federation: 2.3 %Russian Federation: 2.3 %South Africa: 0.1 %South Africa: 0.1 %Sweden: 0.2 %Sweden: 0.2 %Taiwan, China: 0.5 %Taiwan, China: 0.5 %Thailand: 0.2 %Thailand: 0.2 %Turkey: 2.7 %Turkey: 2.7 %Uganda: 0.1 %Uganda: 0.1 %Ukraine: 0.1 %Ukraine: 0.1 %United Kingdom: 0.1 %United Kingdom: 0.1 %United States: 22.9 %United States: 22.9 %Viet Nam: 0.2 %Viet Nam: 0.2 %Zambia: 1.2 %Zambia: 1.2 %Zimbabwe: 0.2 %Zimbabwe: 0.2 %其他BangladeshBelgiumBulgariaCanadaChileChinaCzech RepublicEgyptFranceGermanyGreeceHong Kong, ChinaHungaryIndiaIran (ISLAMIC Republic Of)IraqItalyJapanKazakhstanKorea Republic ofMalaysiaMexicoMongoliaMoroccoNetherlandsNigeriaPeruReservedRomaniaRussian FederationSouth AfricaSwedenTaiwan, ChinaThailandTurkeyUgandaUkraineUnited KingdomUnited StatesViet NamZambiaZimbabwe

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(8)  / Tables(3)

    Share Article

    Article Metrics

    Article Views(1972) PDF Downloads(78) Cited by()
    Proportional views

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return