Next Article in Journal
Novel Approach for Fine Ilmenite Flotation Using Hydrophobized Glass Bubbles as the Buoyant Carrier
Previous Article in Journal
Scale-Up of Decanter Centrifuges for the Particle Separation and Mechanical Dewatering in the Minerals Processing Industry by Means of a Numerical Process Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Leaching of Chalcopyrite under Bacteria–Mineral Contact/Noncontact Leaching Model

1
Technology Center, Shandong Zhaojin Group Co., Ltd., No.108 Shengtai Road, Zhaoyuan 265400, China
2
School of Metallurgy, Northeastern University, 3-11 Wenhua Road, Shenyang 110819, China
*
Author to whom correspondence should be addressed.
Minerals 2021, 11(3), 230; https://doi.org/10.3390/min11030230
Submission received: 16 January 2021 / Revised: 1 February 2021 / Accepted: 4 February 2021 / Published: 24 February 2021
(This article belongs to the Section Mineral Processing and Extractive Metallurgy)

Abstract

:
Bacteria–mineral contact and noncontact leaching models coexist in the bioleaching process. In the present paper, dialysis bags were used to study the bioleaching process by separating the bacteria from the mineral, and the reasons for chalcopyrite surface passivation were discussed. The results show that the copper leaching efficiency of the bacteria–mineral contact model was higher than that of the bacteria–mineral noncontact model. Scanning electron microscopy (SEM), X-ray diffraction (XRD), and Fourier-transform infrared (FTIR) were used to discover that the leaching process led to the formation of a sulfur film to inhibit the diffusion of reactive ions. In addition, the deposited jarosite on chalcopyrite surface was crystallized by the hydrolysis of the excess Fe3+ ions. The depositions passivated the chalcopyrite leaching process. The crystallized jarosite in the bacteria EPS layer belonged to bacteria–mineral contact leaching system, while that in the sulfur films belonged to the bacteria–mineral noncontact system.

1. Introduction

Chalcopyrite (CuFeS2) is one of the most widespread and refractory copper-containing minerals [1]. The main technologies to extract copper are pyrometallurgical processes; however, the byproduct, i.e., sulfur dioxide, leads to serious environmental pollution. Currently, researchers are focusing on developing new environmentally friendly metallurgy technologies and, since bio-metallurgy is renowned for low cost and pollution [2,3], it has already become a research hotspot. There are two widely received bioleaching mechanisms, i.e., direct action and indirect action, providing theoretical support for research on sulfuric mineral bio-metallurgy [4]. On the other hand, further research has shown that these mechanisms cannot explain several complex chemical, electrochemical, and biochemical phenomena. Tributsch found that an extracellular polymeric substance (EPS) is produced as the bacteria are adsorbed onto the ore surface [5]. Fe(Ⅲ) and H+ are accumulated on the contact surface of the ore/solution by EPS to hasten ore decomposition. This conclusion was rather different from the previously characterized “direct action”. Crundwell [6] developed Tributsch’s theory and summarized other research results, eventually establishing the indirect leaching mechanism, indirect contact leaching mechanism, and direct contact leaching mechanism. This mechanism has been already authorized in the bio-metallurgy academic community. In recent years, bio-metallurgy technology developed rapidly and was applied in the smelting process of arsenic-bearing refractory gold ore and secondary copper sulfide [7]. For chalcopyrite bioleaching technology, the development is slow due to the following reasons: the high lattice energy of chalcopyrite [8], its difficulty to be oxidized in the bioleaching process, and the passivation layer produced by the insoluble substance that inhibits the further diffusion of bacteria and reactant [9,10,11,12,13,14,15]. Parker reported that jarosite deposition has an important role in the decrease in copper leaching efficiency [16]. Dutrizac promoted that a dense sulfur layer covered the surface of chalcopyrite which would inhibit leaching [17]. Ghahremaninez promoted that leaching of Cu1−xFe1−yS2 (x < y) was the main reason for the observation of electrochemistry [18,19]. Altogether, the widely accepted view is that jarosite deposition leads to chalcopyrite passivation [20,21].
Fe3+ ion concentration, pH, and monovalent cation concentration have an effect on the generation of the jarosite deposition. Rapid crystallization of jarosite is produced by bacteria, hastening the Fe3+ ion supply [22,23,24,25,26,27]. However, the location of crystallization and mechanism of generation remain unclear. It is very important to study the passivation mechanism of chalcopyrite, especially to ascertain whether jarosite crystallizes on the ore surface or in the EPS layer of bacteria absorbed onto ore surface. In this paper, the chalcopyrite bioleaching processes are divided into two parts by bag filters so as to separate bacteria from ore. The aim was to study the leaching mechanism under different conditions including the bacterium–mineral contact and noncontact leaching models. The leaching residue was analyzed by several experimental methods, such as scanning electron microscopy (SEM), X-ray diffraction (XRD), and Fourier-transform infrared (FTIR) in order to discuss the reason for chalcopyrite surface passivation.

2. Materials and Methods

2.1. Leaching Bacteria

The bioleaching bacterial strains were composed of Acidithiobacillus ferrooxidans, Acidithiobacillus thiooxidans, and Leptospirillum ferriphilum. They were provided by the Northeastern University bio-metallurgical laboratory. The strains have good antitoxic characteristic and oxidative ability of sulfur and ferrous ions, and they grow in a pH range of 1.0–2.5 and a temperature range of 25–44 °C [28,29].

2.2. Microorganism Culture

The culture medium used for this experiment was 9 K [28,29]. The composition of the medium is as follows: 3.0 g/L (NH4)2SO4; 0.1 g/L KCl; 0.5 g/L K2HPO4; 0.5 g/L MgSO4·7H2O; 0.01 g/L Ca(NO3)2; 44.3 g/L FeSO4·7H2O. All chemical reagents used were of analytical grade.

2.3. Mineral Sample

The mineral sample was obtained from a certain copper mine in Hubei province of China. The elementary composition of the sample is listed in Table 1. It can be seen that the mineral sample contained Cu 13.82%, Fe 37.50%, and S 39.76%. The particle size distribution is listed in Table 2. It can be seen that the proportions of particles with size >75 μm, 75–45 μm, 45–37 μm, and <37 μm were 1.12%, 9.53%, 20.16%, and 69.19%, respectively. The X-ray diffraction pattern is shown in Figure 1. The main metallic minerals in the mineral sample were chalcopyrite and pyrite.

2.4. Experimental Procedure

The chalcopyrite samples were divided into two groups for bioleaching. Each group consisted of five parallel samples, as the leaching process needed to be observed and studied under different periods. In the first group, each sample had 10 g of chalcopyrite wrapped with a dialysis bag, which was added to a 500 mL shaking flask containing 180 mL of 9 K culture medium. The dialysis bag was made up of polyvinylidene fluoride with a cutoff relative molecular mass of 8000, preventing bacteria from contacting the mineral but allowing ions to dialyze and diffuse freely. Therefore, the first group test could simulate the leaching behavior of the noncontact bacteria–mineral model, while the second group test could simulate the leaching behavior of the contact bacteria–mineral model. Each sample in the second group had 10 g of chalcopyrite directly added to a 500 mL shaking flask also containing 180 mL of 9 K culture medium. Then, 20 mL of bacterial solution with 3.2 × 108 cells/mL was injected into each of the 10 flasks mentioned above. The initial pH was adjusted to 1.4 with H2SO4. The flasks were put into a constant-temperature shaker incubator and shaken at 170 rpm at 44 °C. The experiment was supplemented with distilled water instead of volatilized water.

2.5. Measurement

A PHS-25 acidometer made by Shanghai digital electronics holding (group) co., LTD was used to measure the solution pH. A platinum plate as a working electrode and a saturated calomel electrode as reference were used to measure redox potential. A Leica DM2500 biological microscope and XB-K-25 hemacytometer were used to count the number of free bacteria in pulp. The total count on the ore surface was measured by ninhydrin colorimetry [30]. The copper concentration was measured using a TAS-990 atomic absorption spectrophotometer, manufactured by PERSEE Company in Beijing. The atomic absorption method is based on the characteristic spectral line of the element to be measured emitted by the hollow cathode lamp. The sample vapor is absorbed by the ground-state atom of the element to be measured in the vapor, and the content of the element to be measured in the sample is determined by the weakening degree of the characteristic spectral line. The leaching residue was observed using an SSX-550 SEM/energy-dispersive spectrometer (EDS) made by the Japanese Shimadzu Company. The phase of leaching residue was analyzed using a Netherlandish D/MAX-RB X-ray diffractometer under the condition of CuKα and 2°/min scanning speed. IR analysis of leaching residue was performed in a 400–4000 cm−1 scanning range by a Vertex70 Fourier-transform infrared spectrometer of the Germany BRUKER Company. Leaching residue was bonded to the metal platform with a conductive adhesive.

3. Results and Discussion

3.1. pH and Potential Change

Figure 2 and Figure 3 show the variations of pH and potential in the process of bioleaching. During the bioleaching process of sulfide minerals, these changes are caused by the transfer of hydrogen ions and electrons. It can be seen from Figure 2 that pH increased at the beginning of the reaction, and then subsequently decreased. Furthermore, the pH remained relatively constant in a range of 1.2–1.75, which is suitable for bacterial survival. Equations (1) and (2) represent the oxidation reaction of sulfide minerals in the bacteria–mineral contact leaching model. Fe3+ as an oxidant comes from two aspects; one is dissociative Fe3+ in the solution, whereas the other occurs in EPS layer of bacteria. The solution pH increases when Fe2+ is oxidized to Fe3+ by the bacteria, according to the reaction shown in Equation (3). However, the pH decreases slowly due to sulfur being oxidized by the bacteria, as shown in Equation (4), and Fe3+ being hydrolyzed, as shown in Equation (5). In the bacteria–mineral noncontact leaching model, the bacteria were isolated from the ore using a dialysis bag, preventing them from being adsorbed onto the ore surface. The chalcopyrite was oxidized only by dissociative Fe3+ in the solution. The standard electrode potential of Fe3+/Fe2+ was 0.771 V vs. SHE which is less than the 0.808 V vs. SHE necessary for the production of S6+ [31,32]. Therefore, S2− in the chalcopyrite was oxidized only to S0, as detected on the ore surface. Fe2+ oxidized to Fe3+ after passing through the dialysis bag and reacting with bacteria, leading to a pH increase. Moreover, S0 on the ore surface could not be oxidized by Fe3+. Thus, the bacteria–mineral noncontact leaching model appeared to have a higher increase in pH than the corresponding contact leaching model. On day 9 of the bioleaching process, the pH reached 1.74, before subsequently decreasing due to Fe3+ hydrolysis, as can be seen in Equation (5). Pulp potential increased rapidly, and then remained relatively constant. Moreover, the potential of the bacteria–mineral noncontact leaching system rose higher than that of the bacteria–mineral contact leaching system. This was due to the sulfur film formed in the bacteria–mineral noncontact leaching system after the ores were oxidized by Fe3+, which in turn obstructed the leaching of ores. Furthermore, the bacteria outside the dialysis bag could oxide Fe2+ into Fe3+, which led to the increase in [Fe3+]/[Fe2+] in the solution, and oxidation reduction potential rose quickly. In the bacteria–mineral contact leaching system, the sulfur film, however, obstructed the leaching of ore to lesser extent due to the existence of Acidithiobacillus thiooxidans. Thus, the sulfide ores could be oxidized continuously [33].
Cu Fe S 2 + 4 F e 3 + chemical C u 2 + + 5 F e 2 + + 2 S 0 .
Fe S 2 + 2 F e 3 + chemical 3 F e 2 + + 2 S 0 .
4 F e 2 + + 4 H + + O 2 biological 4 F e 3 + + 2 H 2 O .
2 S 0 + 3 O 2 + 2 H 2 O biological 2 H 2 S O 4 .
3 F e 3 + + 2 S O 4 2 + 6 H 2 O + M + M F e 3 S O 4 2 OH 6 + 6 H + . ( M = K + ,   NH 4 + ,   H 3 O + )

3.2. Bacterial Concentration Change

The change in bacterial concentration during the process of bioleaching is shown in Figure 4. In the bacteria–mineral contact leaching system, the bacterial concentration is the sum of the adsorbed count and free count. According to Figure 4, leaching from 3 days to 15 days followed the bacterial growth logarithmic phase. Bacterial concentration reached 6 × 108 cells/mL on the 15th day, before approaching the stationary phase. During the whole bioleaching process, the bacterial concentration remained relatively stable. There was no obvious decline phase, indicating that the bacteria adapted very well to the leaching system, and the energy provided by the oxidation of sulfide ore could support the bacteria. In the bacteria–mineral noncontact leaching system, the bacterial concentration increased slowly and reached 2.1 × 108 cells/mL on the 15th day, which was one-third that of the bacteria–mineral contact leaching system. After 15 days, bacterial growth started to decline. The bacterial concentration decreased rapidly and reached 8.2 × 107 cells/mL at the end of the bioleaching process, which shows that Fe3+ had no obvious effect on the oxidation of chalcopyrite in the bacteria–mineral noncontact leaching system. It was so hard for chalcopyrite to decompose persistently that some bacteria struggled to survive. This led to the occurrence of a decline phase and the decrease in bacterial concentration.

3.3. Copper Leaching Efficiency

The copper leaching efficiency during the bioleaching process is shown in Figure 5. As seen from Figure 5, the percentage of leached copper reached 61.67% on the 30th day during the bacteria–mineral contact leaching system, while that of the bacteria–mineral noncontact leaching system reached only 21.86%. In the bacteria–mineral noncontact-leaching system, both the copper leaching rate and the efficiency were very low as a function of the oxidation of dissociative Fe3+. In the bacteria–mineral contact leaching system, the bacteria–EPS–ore system was formed when bacteria were adsorbed on the chalcopyrite surface. Fe3+ was accumulated in the EPS layer, subsequently oxidizing the ore. The efficiency was higher than that of the bacteria–mineral noncontact leaching system. This shows that bacterial adsorption has an important influence on chalcopyrite bioleaching.

3.4. Discussion

3.4.1. XRD Analysis of Leaching Residue

Figure 6 presents the XRD spectrum of several bioleaching residues after chalcopyrite was bioleached for 30 days under two different leaching conditions. As shown in Figure 6, the residue composition of the bacteria–mineral contact leaching system was essentially the same as that of the bacteria–mineral noncontact leaching system, including chalcopyrite, pyrite, elementary sulfur, and jarosite. There were two kinds of leaching mode in the bacteria–mineral contact leaching system. According to the first mode, an EPS layer was formed on chalcopyrite surface by bacteria, and ore was oxidized by Fe3+ enriched in EPS. In the second mode, bacteria, such as Thiobacillus ferrooxidans, were adsorbed onto the ore surface and supported by HS, S0, and S2O32− released by ore decomposition [5,34]. In the earlier stage of chalcopyrite bioleaching, the oxidation rate of ore was faster than that of sulfur by Thiobacillus thiooxidans. Therefore, elementary sulfur accumulated on the chalcopyrite surface. In the bacteria–mineral noncontact leaching system, chalcopyrite was oxidized only by dissociative Fe3+ in the leaching solution, as shown in Equations (1) and (2), such that the sulfur formed could not be oxidized and accumulate on the ore surface. As bioleaching proceeded in the bacteria–mineral contact leaching system, the further oxidation of ore by Acidithiobacillus ferrooxidans was inhibited by sulfur accumulated on the chalcopyrite surface, resulting in a rise in pH in the EPS layer and the hydrolysis of redundant Fe3+. In addition, the jarosite crystal was separated out from the active matter in the EPS layer as a nucleation center. In the bacteria–mineral noncontact leaching system, the chalcopyrite surface was passivated due to the formation of elementary sulfur; thus, chalcopyrite could not effectively be further oxidized by Fe3+. As redundant Fe3+ content increased, Fe3+ was hydrolyzed, and the jarosite crystal was separated out on the chalcopyrite surface. Thus, there was no further increase in the leaching efficiency. Figure 3 shows the XRD spectra of bioleaching residue after bioleaching.

3.4.2. SEM Analysis of Leaching Residue

SEM micrographs of leaching residue at different leaching times are shown in Figure 7 and Figure 8. As shown in Figure 7a, most of the elliptic eroded pits and cracks can be found on the surface of several residues after leaching for 10 days due to bacterial adsorption in the bacteria–mineral contact leaching system. Passivation did not occur due to a slippery ore surface, but fine jarosite crystals existed in the pits. After leaching for 20 days, the pits increased, and the cracks started to broaden and deepen, which clearly indicates that the ores were severely corroded. The ore surface remained sleek, and no jarosite crystal was found on the outer surface of the ore. The jarosite was produced only in the pits. The jarosite appeared only in the eroded area of the ore surface, as shown in Figure 7b. At 20 days through 30 days, since bacteria constantly produced Fe3+, jarosite crystal in the corrosion pits grew, eventually covering the whole chalcopyrite surface, especially where there was no reaction. This led to the thorough passivation of the chalcopyrite surface, as seen in Figure 7a. Figure 8 shows that bacteria were not adsorbed onto the chalcopyrite surface because of the isolation of the bag filter. Chalcopyrite was oxidized by dissociative Fe3+. Elementary sulfur occurred on the chalcopyrite surface after leaching for 10 days. The energy spectrum shows that the crystal floc around sulfur is jarosite, as seen in Figure 8a. Along with the steady oxidation by bacteria, both the Fe3+ concentration and its redox potential increased, resulting in the chemical deposition of jarosite and severe passivation on the chalcopyrite surface. Finally, the copper bioleaching process stopped, as shown in Figure 8b,c.

3.4.3. FTIR Analysis of Leaching Residue

Figure 9 shows the FTIR spectra of both kinds of leaching residue after leaching for 30 days. The adsorption peaks at 515 and 474 cm−1 were attributed to the stretching vibration of the FeO6 octahedron [35], that at 630 cm−1 was attributed to the υ4 stretching vibration of SO42−, that at 1020 cm−1 was attributed to the γ1 bending vibration of SO42−, those at 1090 and 1193 cm−1 were attributed to the γ3 bending vibration of SO42− [36], that at 1663 cm−1 was attributed to the δ deformation vibration of HOH in jarosite, that at 3394 cm−1 was attributed to the υ stretching vibration of OH [27], those at 2958 cm−1 and 2924 cm−1 were attributed to the asymmetric stretching vibration of CH2, and that at 2853 cm−1 was attributed to the symmetrical stretching vibration of CH2 [23,37]. The figure shows that the bacteria–mineral contact leaching system possessed the same structure of jarosite as the bacteria–mineral noncontact leaching system. Organic matter absorbed in the bacteria–mineral contact leaching system was derived from the deposition of biomass on the chalcopyrite surface. The biomass deposition hastened the nucleation and crystallization of jarosite.

3.4.4. Model of Bioleaching Process

Although bacteria–mineral contact and noncontact leaching coexist in chalcopyrite bioleaching, bacteria–mineral contact leaching represents the main mode. In Figure 10, a mechanism model is presented for the chalcopyrite bioleaching process and jarosite production. In the initial phase, bacteria are adsorbed onto the mineral surface owing to its chemotaxis, as shown in Figure 10a. A substantial amount of Fe3+ exists in the EPS layer, which is reduced to Fe2+ due to the oxidation of ore. Then, Fe2+ is oxidized to Fe3+, which is accompanied extensive H+ consumption, as seen in Figure 10b,c. Chalcopyrite is oxidized, and Fe and Cu are liberated. The sulfur in the reduced state is oxidized to S0, and a sulfur film is formed on the chalcopyrite surface, thereby inhibiting the leaching of chalcopyrite. The copper leaching rate decreases and pH increases in the EPS layer. The remaining Fe3+ hydrolyzes and forms jarosite, as shown in Figure 10d. In the EPS layer, active matter is used as the nucleation center of jarosite. Fe2+ is oxidized by free bacteria, thereby increasing the Fe3+ concentration and redox potential. During the anaphase, the jarosite sediment is increased, thereby inhibiting the copper leaching process with the rise in Fe3+ concentration. Thus, passivation of the bacteria–mineral contact and noncontact leaching systems depends on the accumulation of sulfur on the chalcopyrite surface. In the passivation process, the formation of a sulfur film has a inhibitory effect on the rapid diffusion of reaction ions, thus decreasing the leaching rate of chalcopyrite. However, jarosite is the main reason for chalcopyrite passivation.

4. Conclusions

(1) The bacteria–mineral contact and noncontact leaching systems coexisted in the chalcopyrite bioleaching process. In the bacteria–mineral contact leaching system, the bacteria oxidized elemental sulfur on the surface of chalcopyrite and enriched Fe3+ in the EPS layer, which hastened copper leaching. In the bacteria–mineral noncontact leaching system, dissociative Fe3+ oxidized chalcopyrite, and it was regenerated by the bacteria.
(2) In the chalcopyrite bioleaching process, the formation of sulfur was found in both the bacteria–mineral contact and the bacteria–mineral noncontact leaching systems. In the bacteria–mineral contact leaching system, the oxidation rate of ore was faster than that of sulfur by Thiobacillus thiooxidans. Sulfur was deposited on the chalcopyrite surface. In the bacteria–mineral noncontact leaching system, the oxidation of S2−and S1− to S0 induced by Fe3+ could not generate S4+ and S6+ with a higher valence state, consequently inhibiting further copper leaching.
(3) A layer of sulfur was produced and consequently inhibited the rapid diffusion of reaction ions. The redundant Fe3+ was hydrolyzed and formed jarosite. In the bacteria–mineral contact leaching system, jarosite was separated out from the active matter in the EPS layer as a nucleation center. In the bacteria–mineral noncontact leaching system, the jarosite crystallized on the sulfur layer of the chalcopyrite surface.

Author Contributions

Conceptualization, P.M.; methodology, P.M. and H.Y.; validation, H.Y.; formal analysis, P.M.; investigation, P.M. and Z.L.; data curation, P.M.; writing—original draft preparation, P.M.; writing—review and editing, P.M., Q.S., A.A. and L.T.; funding acquisition, Z.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key R&D Program of China (2018YFC1902003).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The study did not report any data.

Acknowledgments

We greatly thank all anonymous reviewers for their constructive comments, which improved the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Brierley, J.A. Acdidophilic thermophilic archaebacteria: Potential application for metals recovery. FEMS Microbiol. Lett. 1990, 75, 287–291. [Google Scholar] [CrossRef]
  2. Petersen, J.; Dixon, D.G. Competitive bioleaching of pyrite and chalcopyrite. Hydrometallurgy 2006, 83, 40–49. [Google Scholar] [CrossRef]
  3. Brierley, J.A.; Brierley, C.L. Present and future commercial applications of biohydrometallurgy. Hydrometallurgy 2001, 59, 233–239. [Google Scholar] [CrossRef]
  4. Silverman, M.P.; Ehrlich, H.L. Microbial formation and degradation of minerals. Adv. Appl. Microbiol. 1964, 6, 153–206. [Google Scholar]
  5. Tributsch, H. Direct versus indirect bioleaching. Hydrometallurgy 2001, 59, 177–185. [Google Scholar] [CrossRef]
  6. Crundwell, F.K. How do bacteria interact with minerals? Hydrometallurgy 2003, 71, 75–81. [Google Scholar] [CrossRef]
  7. Brierley, C.L. Biohydrometallurgical prospects. Hydrometallurgy 2010, 104, 324–328. [Google Scholar] [CrossRef]
  8. Hiroyoshi, N.; Kitagawa, H.; Tsunekawa, M. Effect of solution composition on the optimum redox potential for chalcopyrite leaching in sulfuric acid solutions. Hydrometallurgy 2008, 91, 144–149. [Google Scholar] [CrossRef]
  9. Klauber, C. A critical review of the surface chemistry of acidic ferric sulphate dissolution of chalcopyrite with regards to hindered dissolution. Int. J. Miner. Process. 2008, 86, 1–17. [Google Scholar] [CrossRef]
  10. Kinnunen, H.M.; Heimala, S.; Riekkola-Vanhanen, M.L.; Puhakka, J.A. Chalcopyrite concentrate leaching with biologically produced ferric sulphate. Bioresour. Technol. 2006, 97, 1727–1734. [Google Scholar] [CrossRef]
  11. Rawlings, D.E.; Johnson, B.D. Biomining; Springer: Berlin/Heidelberg, Germany, 2007; pp. 287–301. [Google Scholar]
  12. Yohana, R.; Ballester, A.; María, L.B.; Felisa, G.; Jesús, A.M. Study of bacterial attachment during the bioleaching of pyrite, chalcopyrite, and sphalerite. Geomicrobiology 2003, 20, 131–141. [Google Scholar]
  13. Yang, Y.; Harmer, S.; Chen, M. Synchrotron X-ray photoelectron spectroscopic study of the chalcopyrite leached by moderate thermophiles and mesophiles. Miner. Eng. 2014, 69, 185–195. [Google Scholar] [CrossRef]
  14. Akcil, A.; Ciftci, H.; Deveci, H. Role and contribution of pure and mixed cultures of mesophiles in bioleaching of a pyritic chalcopyrite concentrate. Miner. Eng. 2007, 20, 310–318. [Google Scholar] [CrossRef]
  15. Ahmadi, A.; Schaffie, M.; Manafi, Z.; Ranjbar, M. Electrochemical bioleaching of high grade chalcopyrite flotation concentrates in a stirred bioreactor. Hydrometallurgy 2010, 104, 99–105. [Google Scholar] [CrossRef]
  16. Parker, A.; Klauber, C.; Kougianos, A.; Watling, H.R.; Van-Bronswijk, W. An X-ray photoelectron spectroscopy study of the mechanism of oxidative dissolution of chalcopyrite. Hydrometallurgy 2003, 71, 265–276. [Google Scholar] [CrossRef]
  17. Dutrizac, J.E. Elemental sulphur formation during the ferric chloride leaching of chalcopyrite. Hydrometallurgy 1990, 23, 153–176. [Google Scholar] [CrossRef]
  18. Ghahremaninezhad, A.; Dixon, D.G.; Asselin, E. Kinetics of the ferric–ferrous couple on anodically passivated chalcopyrite (CuFeS2) electrodes. Hydrometallurgy 2012, 125–126, 42–44. [Google Scholar] [CrossRef]
  19. Yao, G.; Wen, J.; Gao, H.; Wang, D. Chalcopyrite bioleaching by moderate thermophilic bacteria and surface passivation. J. Cent. South Univ. (Sci. Technol.) 2010, 41, 1234–1236. (In Chinese) [Google Scholar]
  20. Pan, H.; Yang, H.; Tong, L.; Zhong, C.; Zhao, Y. Control method of chalcopyrite passivation in bioleaching. Trans. Nonferrous Met. Soc. China 2012, 22, 2255–2260. (In Chinese) [Google Scholar] [CrossRef]
  21. Yang, H.; Pan, H.; Tong, L.; Liu, Y. Formation process of biological oxide film on chalcopyrite crystal surface. Acta Metall. Sin. 2012, 48, 1145–1152. (In Chinese) [Google Scholar] [CrossRef]
  22. Lu, X.; Lu, J.; Zhu, C.; Liu, X.; Wang, R.; Li, Q.; Xu, Z. Preliminary study on surface properties of iron sulfate formed by microbially Induced mineralization. Geol. J. China Univ. 2005, 2, 194–198. (In Chinese) [Google Scholar]
  23. Henao, D.M.O.; Godoy, M.A.M. Jarosite pseudomorph formation from arsenopyrite oxidation using Acidithiobacillus ferrooxidans. Hydrometallurgy 2010, 104, 162–168. [Google Scholar] [CrossRef]
  24. Zhu, C.; Lu, J.; Lu, X.; Wang, R.; Li, Q. SEM study on jarosite mediated by thiobacillus ferrooxidans. Geol. J. China Univ. 2005, 11, 234–238. (In Chinese) [Google Scholar]
  25. Bai, S.; Liang, J.; Wang, M.; Zhou, L. Effect of Fe(III) supply rate on the formation of amorphous schwertmannite. Acta Mineral. Sin. 2011, 2, 256–262. (In Chinese) [Google Scholar]
  26. Daoud, J.; Karamanev, D. Formation of jarosite during Fe2+ oxidation by Acidithiobacillus ferrooxidans. Miner. Eng. 2006, 19, 960–967. [Google Scholar] [CrossRef]
  27. Sasaki, K.; Konno, H. Morphology of jarosite-group compounds precipitated from biologically and chemically oxidized Fe ions. Can. Mineral. 2000, 38, 45–56. [Google Scholar] [CrossRef]
  28. Fang, D.; Zhou, L.X. Effect of sludge dissolved organic matter on oxidation of ferrous iron and sulfur by Acidithiobacillus ferrooxidans and Acidithiobacillus thiooxidans. Water Air Soil Pollut. 2006, 171, 81–94. [Google Scholar] [CrossRef]
  29. Li, J.; Tong, L.; Xia, Y.; Yang, H.; Auwalu, A. Microbial synergy and stoichiometry in heap biooxidation of low-grade porphyry arsenic-bearing gold ore. Extremophiles 2020, 24, 355–364. [Google Scholar] [CrossRef]
  30. Pan, H.; Yang, H.; Chen, S. Conditions for biosorption measuring on chalcopyrite surface. J. Northeast. Univ. (Nat. Sci.) 2010, 31, 999–1002. (In Chinese) [Google Scholar]
  31. Bevilaqua, D.; Lette, A.L.L.C.; Garcia, J.O.; Tuovinen, O.H. Oxidation of chalcopyrite by Acidithiobacillus ferrooxidans and Acidithiobacillus thiooxidans in shake flasks. Process Biochem. 2002, 38, 587–592. [Google Scholar] [CrossRef]
  32. Ghahremaninezhad, A.; Asselin, E.; Dixon, D.G. Electrochemical evaluation of the surface of chalcopyrite during dissolution in sulfuric acid solution. Electrochim. Acta 2010, 55, 5041–5056. [Google Scholar] [CrossRef]
  33. Klauber, C.; Parker, A.; Van-Bronswijk, W.; Watling, H. Sulphur speciation of leached chalcopyrite surfaces as determined by X-ray photoelectron spectroscopy. Int. J. Miner. Process. 2001, 62, 65–94. [Google Scholar] [CrossRef]
  34. Sand, W.; Gehrke, T. Extracellular polymeric substances mediate bioleaching/biocorrosion via interfacial processes involving iron(III) ions and acidophilic bacteria. Res. Microbiol. 2006, 157, 49–56. [Google Scholar] [CrossRef]
  35. Powers, D.A.; Rossman, G.R.; Schugar, H.J.; Gray, H.B. Magnetic behavior and infrared spectra of jarosite, basic iron sulfate, and their chromate analogs. J. Solid State Chem. 1975, 13, 1–13. [Google Scholar] [CrossRef]
  36. Lazaroff, N.; Sigal, W.; Wasserman, A. Iron oxidation and precipitation of ferric hydroxysulfates by resting Thiobacillus ferrooxidans cells. Appl. Environ. Microbiol. 1982, 43, 924–938. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Marel, H.W.; Beutelspacher, H. Atlas of Infrared Spectroscopy of Clay Minerals and Their Admixtures; Elsevier Publishing Company: Amsterdam, The Netherlands, 1976; p. 396. [Google Scholar]
Figure 1. X-ray diffraction (XRD) pattern of the mineral samples.
Figure 1. X-ray diffraction (XRD) pattern of the mineral samples.
Minerals 11 00230 g001
Figure 2. pH changes with bioleaching time.
Figure 2. pH changes with bioleaching time.
Minerals 11 00230 g002
Figure 3. Eh changes with bioleaching time.
Figure 3. Eh changes with bioleaching time.
Minerals 11 00230 g003
Figure 4. Bacterial concentration changes with bioleaching time.
Figure 4. Bacterial concentration changes with bioleaching time.
Minerals 11 00230 g004
Figure 5. Copper leaching efficiency with bioleaching time.
Figure 5. Copper leaching efficiency with bioleaching time.
Minerals 11 00230 g005
Figure 6. XRD spectra of the leaching residue after 30 days.
Figure 6. XRD spectra of the leaching residue after 30 days.
Minerals 11 00230 g006
Figure 7. The morphologies of leaching residue after (a) 10 days, (b) 20 days, and (c) 30 days in the bacteria–mineral contact leaching system (SE).
Figure 7. The morphologies of leaching residue after (a) 10 days, (b) 20 days, and (c) 30 days in the bacteria–mineral contact leaching system (SE).
Minerals 11 00230 g007
Figure 8. The morphologies of leaching residue after (a) 10 days, (b) 20 days, and (c) 30 days in the bacteria–mineral noncontact leaching system (SE).
Figure 8. The morphologies of leaching residue after (a) 10 days, (b) 20 days, and (c) 30 days in the bacteria–mineral noncontact leaching system (SE).
Minerals 11 00230 g008
Figure 9. Fourier-transform infrared (FTIR) spectra of leaching residue.
Figure 9. Fourier-transform infrared (FTIR) spectra of leaching residue.
Minerals 11 00230 g009
Figure 10. Schematic diagram of chalcopyrite bioleaching and jarosite formation. (a) Adsorption bacteria on chalcopyrite surface, (b) Microenvironment between bacteria, EPS and minerals, (c) Fe2+ was oxidized to Fe3+ on the cell wall of bacteria, and consuming H+, (d) Jarosite crystal precipitation.
Figure 10. Schematic diagram of chalcopyrite bioleaching and jarosite formation. (a) Adsorption bacteria on chalcopyrite surface, (b) Microenvironment between bacteria, EPS and minerals, (c) Fe2+ was oxidized to Fe3+ on the cell wall of bacteria, and consuming H+, (d) Jarosite crystal precipitation.
Minerals 11 00230 g010
Table 1. The main element contents of mineral samples.
Table 1. The main element contents of mineral samples.
ElementCuFeSSiO2CaOAl2O3MgO
Content, %13.8237.5039.762.720.790.400.15
Table 2. Particle size distribution of mineral samples.
Table 2. Particle size distribution of mineral samples.
Particle Size>75 μm75–45 μm45–37 μm<37 μm
Content, %1.129.5320.1669.19
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ma, P.; Yang, H.; Luan, Z.; Sun, Q.; Ali, A.; Tong, L. Leaching of Chalcopyrite under Bacteria–Mineral Contact/Noncontact Leaching Model. Minerals 2021, 11, 230. https://doi.org/10.3390/min11030230

AMA Style

Ma P, Yang H, Luan Z, Sun Q, Ali A, Tong L. Leaching of Chalcopyrite under Bacteria–Mineral Contact/Noncontact Leaching Model. Minerals. 2021; 11(3):230. https://doi.org/10.3390/min11030230

Chicago/Turabian Style

Ma, Pengcheng, Hongying Yang, Zuochun Luan, Qifei Sun, Auwalu Ali, and Linlin Tong. 2021. "Leaching of Chalcopyrite under Bacteria–Mineral Contact/Noncontact Leaching Model" Minerals 11, no. 3: 230. https://doi.org/10.3390/min11030230

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop