Next Article in Journal
Electrochemical Polishing of Ti6Al4V Alloy Assisted by High-Speed Flow of Micro-Abrasive Particles in NaNO3 Electrolyte
Previous Article in Journal
Precipitation Law of Vanadium in Microalloyed Steel and Its Performance Influencing Factors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthetic Sulfide Concentrate Dissolution Kinetics in HNO3 Media

Laboratory of Advanced Technologies in Non-Ferrous and Ferrous Metals Raw Materials Processing, Department of Non-Ferrous Metals Metallurgy, Ural Federal University, 620002 Yekaterinburg, Russia
*
Author to whom correspondence should be addressed.
Materials 2022, 15(22), 8149; https://doi.org/10.3390/ma15228149
Submission received: 13 October 2022 / Revised: 10 November 2022 / Accepted: 14 November 2022 / Published: 17 November 2022

Abstract

:
The nature of tennantite (Cu12As4S13), chalcopyrite (CuFeS2) and sphalerite (ZnS) particles’ mixture dissolution in nitric acid (HNO3) media was investigated in this study. The effects of temperature (323–368 K), HNO3 (1–8 mol/L) and Fe3+ (0.009–0.036 mol/L) concentrations, reaction time (0–60 min) and pyrite (FeS2) additive (0.5/1–2/1; FeS2/sulf.conc.) on the conversion of the minerals were evaluated. It has been experimentally shown that the dissolution of the mixture under optimal conditions (>353 K; 6 mol/L HNO3; FeS2/synt. conc = 1/1) allows Cu12As4S13, CuFeS2 and ZnS conversion to exceed 90%. The shrinking core model (SCM) was applied for describing the kinetics of the conversion processes. The values of Ea were calculated as 28.8, 33.7 and 53.7 kJ/mol, respectively, for Cu12As4S13, CuFeS2 and ZnS. Orders of the reactions with respect to each reactant were calculated and the kinetic equations were derived to describe the dissolution rate of the minerals. It was found that the interaction between HNO3 solution and Cu12As4S13, CuFeS2 and ZnS under the conditions investigated in this are of a diffusion-controlled nature. Additionally, the roles of Fe(III) in the initial solution and FeS2 in the initial pulp as catalysts were studied. The results indicated that the increase in Fe3+ concentration significantly accelerates the dissolution of the mixture, while the addition of FeS2 forms a galvanic coupling between FeS2, and Cu12As4S13 and CuFeS2, which also accelerates the reaction rate. The results of the study are considered useful in developing a hydrometallurgical process for polymetallic sulfide raw materials treatment.

1. Introduction

Intensive exploitation of the primary raw materials in the non-ferrous metals industry resulted in the depletion of the rich deposits as a consequence of seeking alternative sources of raw materials. In view of the current trend in decreasing the metal content in the ore, smelters adapt to lower grade and technogenic raw materials [1]. One of the most specific raw material types are arsenic-containing ores and concentrates since their treatment is associated with additional environmental risks and technological complexity. Thus, along with the familiar minerals in copper industry such as chalcopyrite (CuFeS2), covellite (CuS), chalcocite (Cu2S), bornite (Cu5FeS4) and sphalerite (ZnS), sulfide ores occasionally contain minerals of the Fachlor group, such as tennantite (Cu12As4S13) and tetrahedrite (Cu12Sb4S13), which complicates the processing of the obtained ore/concentrates using typical approaches [2].
Pyrometallurgical processing of the arsenic-containing raw materials admits releasing arsenic dust and other waste into the atmosphere, which poses a significant danger to humans and the environment [3,4,5,6,7,8,9,10,11,12]. Therefore, pyrometallurgical processing of such materials is considered unacceptable today, and hence, undergoes strict government regulation in the most countries.
As an alternative to conventional smelting, various hydrometallurgical approaches are currently being researched, including the following: acid [13,14,15] and ammonia leaching [16], alkaline [17,18,19] and autoclave approaches [20,21,22], bioleaching [23,24], etc. Although autoclave approaches have been known since the middle of the last century and have found success and wide application for gold–silver sulfide raw materials treatment [25], atmospheric systems are also of great interest in hydrometallurgy [26].
One of the most promising approaches in this direction is an application of nitric acid (HNO3) as a reactant for atmospheric leaching, since HNO3 allows the most complete decomposition of the sulfide matrix [27,28]. Thus, HNO3 leaching under optimal conditions could provide both the transference of valuable metals into solution and arsenic and other toxic elements into stable and environmentally friendly compounds [29,30,31,32,33].
This study is aimed at evaluating the fundamental effectiveness of sulfide low-grade raw-materials treatment using HNO3 as the reactant for atmospheric leaching and also at evaluating the kinetic characteristics of the dissolution processes. High-purity specimens of Cu12As4S13, CuFeS2 and ZnS were used in this study to create the mixture of sulfides (synthetic concentrate), which was subjected dissolution in HNO3 media. The choice of these minerals was for the purpose of creating the most refractory raw materials through hydrometallurgical treatment.
The shrinking core model (SCM) was applied for describing the kinetics of the dissolution processes. The effects of temperature, reaction time and HNO3 concentration on the conversion of the minerals were evaluated. We also evaluated the effect of the pyrite (FeS2) additive and the Fe3+ concentration in the initial solution, since a number of studies reported [34,35,36,37,38,39,40,41,42,43,44] using the mentioned reactants as catalysts.
The results of the study are considered useful in developing hydrometallurgical processes for polymetallic sulfide raw-materials treatment.

2. Materials and Methods

2.1. Materials

High-purity specimens of Cu12As4S13, CuFeS2, ZnS and FeS2 originating from the Uchalinskii deposit (Sverdlovsk region, Russia), the Vorontsovskii deposit (North Ural region, Russia), the Karabashskii deposit (South Ural region, Russia) and the Berezovskii deposit (Sverdlovsk region, Russia), respectively, were used in this study (Figure 1A–D). The samples for experiments were obtained from the ground crystals by wet sieving (20–40 µm), following the mixing of Cu12As4S13, CuFeS2 and ZnS in the proportion of 1/0.36/0.17 (by weight). The obtained mixture represented the sulfide concentrate of the complex composition (Table 1). The established ratio of the minerals in synthetic concentrate is typical for the industrial Cu-As concentrate of the Uchalinskii deposit [44]. An additive of FeS2 into the leaching pulp was used in the study to evaluate the galvanic effect between FeS2, Cu12As4S13 and CuFeS2.

2.2. Experimental Procedure

The dissolution of the synthetic concentrate in HNO3 solution was conducted in a 500 mL round-bottom borosilicate glass reaction vessel (Lenz Laborglas GmbH & Co. KG, Wertheim, Germany) with a thermostatic jacket, which was thermostated using a Huber KISS-205B circulator (Huber Kältemaschinenbau AG, Offenburg, Germany). The reactor is equipped with an IKA EUROSTAR 20 digital overhead stirrer (IKA®-Werke GmbH & Co. KG, Staufen, Germany).
The values of the change in the Gibbs energy were calculated using the HSC Chemistry Software v. 9.9 (Metso Outotec Finland Oy, Tampere, Finland).
To examine the synthetic concentrate dissolution in HNO3 solution, a 5 g synthetic concentrate sample (particle size +20–40 µm) was added into the 200 mL of 1–8 mol/L HNO3 solution when the temperature inside the leaching vessel reached the desired temperature (323–368 K). When applicable, an additive of FeS2 was added to the initial pulp with different mass ratios of FeS2 to synthetic concentrate sample: 0.5/1, 1/1, 1.5/1, 2/1. The latter was equal to 2.5, 5, 7.5 and 10 g of FeS2 additive, respectively. During the experimental runs, samples were withdrawn from the reaction vessel at regular time intervals (1, 1.5, 2, 5, 15, 30, 45 and 60 min) using an automatic dispenser Sartorius Proline (Mine-beaIntecAachen GmbH & Co. KG, Aachen, Germany) and filtered using a 45 µm syringe filter. The final pulp after the experiments was filtered using a Buchner funnel with filter paper. The solid residue (cake) was washed with distilled water, dried in an oven at a temperature of 274 K at least for 1 h, weighed, and analyzed further for the residue characterization. The solution samples received during the experiments as well as solutions after filtration of the final pulp were subjected to volume measuring and analysis for Cu, Fe, As and Zn. The received data was used in calculating the fraction reacted (X) and the conversion (E, %) of the minerals with the following equations:
X = m s m i
E =   m s m i · 100 %
where ms and mi are the mass of As, Cu and Zn in solution after the treatment and the mass of As, Cu, Zn in the initial synthetic concentrate, respectively.

2.3. Analysis

The mineralogical and chemical compositions of the minerals and solid residues were determined based on wave dispersive spectrometry (ARL Advant’X 4200, Thermo Fisher Scientific, Walthamm, MA, USA), X-ray diffraction “XRD” (XRD-7000, Shimadzu, Kyoto, Japan), scanning electron microscopy “SEM” (JSM-6390LV, Jeol, Tokyo, Japan) with a JED 2300 Energy Dispersive X-ray Analyzer (EDX) (Jeol, Tokyo, Japan), wet analysis using inductively coupled plasma mass-spectrometry “ICP-MS” (Elan 9000, PerkinElmer, Waltham, MA, USA) and laser diffraction (Helos/BR, Sympatec, Clausthal-Zellerfeld, Germany). The solid materials were ground in a planetary mill (Pulverisette 6, Fritsch GmbH & Co. KG, Welden Germany) and dissolved in an aqua regia before ICP-MS analysis. The sulfur content was analyzed using a carbon/sulfur analyzer (CS 230, LECO, St. Joseph, MI, USA). Solution samples were analyzed by ICP-MS.

3. Results and Discussion

3.1. Effect of Temperature

Temperature range under investigation in this study was taken as 323–368 K, since previously published works [45] show a low conversion of sulfide minerals at ambient temperature. Thus, the chosen temperature range is of primary interest for the atmospheric leaching processes investigation.
Increasing the temperature seemed to have a significant effect on the conversion of Cu12As4S13 (Figure 2a), CuFeS2 (Figure 2b) and ZnS (Figure 2c). To illustrate, at 323 K for 60 min, only 65% of Cu12As4S13, 45% of CuFeS2 and 53% of ZnS were reacted, while at 368 K conversion increased to 96, 75 and 98%, respectively for Cu12As4S13, CuFeS2 and ZnS.
According to the conversion profiles, the progress of the reaction was observed to slow down with time. Additionally, the presence of a clear inflection line in curves allowed us to suggest an internal diffusion mechanism of the interactions. The possible internal diffusion may cause elemental sulfur formation during the processing of the synthetic concentrate by the following reactions:
CuFeS2 + 16HNO3 = FeSO4 + CuSO4 + 16NO2+ 8H2O; ΔG0353 = −1187 kJ/mol,
CuFeS2 + 10HNO3 = Fe(NO3)3 + Cu(NO3)2 +2S0 + 5NO2 + 5H2O; ΔG0353 = −438 kJ/mol,
Cu12As4S13 + 64HNO3 = 12Cu(NO3)2 + 4H3AsO4 + 13H2SO4 + 40NO + 13H2O; ΔG0353 = −1866 kJ/mol,
Cu12As4S13 + 38HNO3 = 12Cu(NO3)2 + 4H3AsO4 + 13S0 + 14NO + 13H2O; ΔG0353 = −762 kJ/mol,
ZnS + 8HNO3 = ZnSO4 + H2SO4 + 8NO2 + 4H2O; ΔG0353 = −640 kJ/mol

3.2. Effect of HNO3 Concentration

Figure 3 illustrates the effect of increasing the initial concentration of HNO3 ranging from 1 to 8 mol/L on the conversion of Cu12As4S13, CuFeS2 and ZnS. The increase in HNO3 concentration seemed to also improve the rate and extent of the sulfides conversion. Over the reaction period, the conversion of Cu12As4S13 (Figure 3a), CuFeS2 (Figure 3b) and ZnS (Figure 3c) significantly increased from 35, 24 and 67%, respectively, at 1 mol/L HNO3 in solution to 94, 81 and 98% at 8 mol/L HNO3.
The significant effect of HNO3 concentration on the reaction rate and conversion extent may also indicate that the reactions are controlled by diffusion through the product layer, where the increasing HNO3 concentration in the initial solution leads to acceleration of S° to SO42− transformation and as consequence, the predominant course of reactions by Equations (3), (5) and (7).

3.3. Effect of Fe(III) Concentration

The effect of Fe (III) concentration ranging from 0.009 to 0.036 mol/L on the conversion of Cu12As4S13, CuFeS2 and ZnS was investigated. The results in Figure 4 show a moderate increase in the reaction rate by increasing the Fe (III) concentration. After 60 min of reaction at 0.009 mol/L Fe (III), 79, 65 and 91% of Cu12As4S13 (Figure 4a), CuFeS2 (Figure 4b) and ZnS (Figure 4c), respectively, were converted, compared with 90, 96 and 98% conversions, respectively at 0.036 mol/L. The interaction of main minerals in synthetic concentrate with Fe(III) are supposed to proceed according to the following reactions:
CuFeS2 + 2Fe2(SO4)3 = CuSO4 + 5FeSO4 + 2S0; ΔG0353 = −65 kJ/mol,
ZnS + Fe2(SO4)3 = ZnSO4 + 2FeSO4 + S0; ΔG0353 = −56 kJ/mol
Cu12As4S13 + 13.5Fe2(SO4)3 + 6H2O = 12CuSO4 + 27FeSO4 + 14.5S0 + 4H3AsO3; ΔG0353 = −28 kJ/mol

3.4. Effect of FeS2 Additive

Four different mass ratios of FeS2 to synthetic concentrate (0.5/1, 1/1, 1.5/1, 2/1, which is equal to 2.5, 5, 7.5 and 10 g of FeS2 adding) were used in the experiments to examine the effect of galvanic coupling. The results are shown in Figure 5.
As expected, increasing the mass of additive resulted in moderately improved conversion [35]. The obtained results indicate that the oxidation of Cu12As4S13 and CuFeS2 (Equations (4) and (6)) may proceed with the formation of S°, which passivates the surface of the minerals. At the same time, FeS2 may act as an alternative catalytic surface for these minerals. The latter provides the reduction of HNO3 on FeS2 surface and decomposition of other sulfide minerals [45].
In the experiments with 0.5/1 mass ratio, the conversion of Cu12As4S13 (Figure 5a), CuFeS2 (Figure 5b) and ZnS (Figure 5c) appeared to be limited to 64, 58 and 89%, respectively. Usage of the higher mass ratio (2/1) resulting the conversion of the minerals to be increased up to 83, 83 and 98%, respectively for Cu12As4S13, CuFeS2 and ZnS.

4. Characterization of Residues

Characteristics of the Received Cakes

Figure 6 shows the SEM images (Figure 6a,b) and EDX-mapping of the residue (Figure 6c–f) after leaching the synthetic concentrate in HNO3 solution. EDX-mapping images confirm the formation of S° layer on the surface of unreacted synthetic concentrate particles. Thus, the dark layer over the muted points of Fe (Figure 6c) and Cu (Figure 6d) as well as bright green points on the image of joint EDX-mapping (Figure 6f) over all other components allow us to confirm the S° presence. Thus, the mentioned conditions suggest proceeding interactions by Equations (4) and (6).
The S° content in the residue was observed to be 56% and the conversion of sulfide sulfur to S° appeared to be about 38%. Under these conditions, the conversion of Cu12As4S13, CuFeS2, and ZnS was 59, 60 and 84%, respectively.
In contrast to the experiment without the FeS2 additive (Figure 6), the results with the additive (FeS2/synt. conc = 1/1) showed a lesser S° content.
The Fe and Cu points indicated in one component EDX-mapping images (Figure 7c,d, respectively) became brighter. According to Figure 7a,b, the residue has a heterogeneous surface resembling conglomerates, while the experiment without FeS2 additive shows a more homogeneous structure due to the covering of the particles by S°. The green zones in Figure 7f correspond to the distribution of S°, while the mixture of red and blue zones are copper minerals (Cu12As4S13 and CuFeS2) and FeS2.
Therefore, S° covers the surface of the synthetic concentrate in lesser extent, which confirms by the SEM-EDX residue investigation as well as the chemical composition of the residue—sulfide sulfur to S° transformation decreased to 23%, while the S° content in the solid residue decreased to 14%. Under these conditions of dissolution, the conversion of Cu12As4S13, CuFeS2 and ZnS was 87, 91 and 98%, respectively.
Figure 8 shows the XRD patterns of the solid residues after the dissolution of synthetic concentrate in HNO3 solution. The obtained data additionally confirms that the presence of FeS2 allows to limit the formation of S°.
SEM images (Figure 9) of the material after dissolution for 15 min (bend point in Figure 5 for Cu12As4S13 and CuFeS2) coupled with EDX analysis (Table 2) suggest that most of the S° was formed towards the end of this period and the subsequent dissolution of the material occurs at its coating by S°.
Therefore, it is appropriate to conclude that the diffusion in the system is the result of S° formation during the first 10–20 min of the experiment. After that, the dissolution process shifts to diffusion control.

5. Kinetics Analysis

As it was shown, the conversion of sulfides is significantly affected by temperature, HNO3 concentration and the presence of FeS2 in the system; that could mean possible control of the reactions by both chemical reaction and diffusion. To determine the limiting stage of the processes, the most commonly used kinetic equations of SCM describing liquid-solid reactions [46] were used to fit into the experimental data (Table 3). According to the results present in Figure 2, Figure 3, Figure 4 and Figure 5, the higher conversion degree during the initial period of reaction was observed for ZnS, therefore, kinetic analysis of the mineral was carried out in the period from 0 to 2 min, while for Cu12As4S13 and CuFeS2, from 0 to 60 min.
As shown in Figure 10, the SCM equation typically applied for diffusion kinetic system (Table 4, Equation (1)) can be used to describe the conversion processes with high values of the determination coefficient (R2).
The activation energy values (Ea) were calculated using the Arrhenius law (Figure 11). Thus, Ea was determined as 28.8 kJ/mol for Cu12As4S13 and 33.7 kJ/mol for CuFeS2, values that are typical for inner-diffusion processes [46]. The activation energy values for ZnS treatment were determined as 53.7 kJ/mol, which is more typical for kinetically controlled processes. However, according to the literature [47,48,49,50,51,52], a high Ea value is not always allowed to make the final decision on the process nature.
The reaction order with respect to HNO3 concentration, Fe (III) ions concentration and amount of FeS2 additive were calculated using the graphical method (Table 4). The fractional order of reaction with respect to Fe (III) ions concentration and amount of FeS2 additive at Cu12As4S13, CuFeS2 and ZnS treatment suggests that the nature of the processes is diffusion controlled. At the same time, the reaction order with respect to HNO3 concentration at Cu12As4S13, CuFeS2 and ZnS treatment is more typical for chemical reaction control. The latter could be a result of aggressive impact on the S° layer that allows it to overcome the effect of passivation.
As a result, the research data were generalized and the general kinetic equations were established separately for Cu12As4S13, CuFeS2 and ZnS treatment, which consider the influence of temperature, concentration of reagents and duration of the experiments. As it shown in Figure 12, the relationship between the equations 1 – 3(1 – X)2/3 + 2·(1 – X) and CHNO3·CFe(III)·CFeS2·exp[–Ea/(R·T)]·τ·103 for all experimental data was established, and the data points were evenly distributed along straight lines with a high R2.
The kinetic equations for treatment of Cu12As4S13, CuFeS2 and ZnS can be written as follows (11)–(13), respectively:
Cu12As4S13: 1 − 3(1 − X) 2/3 + 2(1 − X) = 38820CHNO31.2CFe(III)0.34CFeS20.47e−28858/RTt
CuFeS2: 1 − 3(1 − X) 2/3 + 2(1 − X) = 74070CHNO31.42CFe(III)0.82CFeS20.69e−33708/RTt
ZnS:1 − 3(1 − X) 2/3 + 2 (1 − X) = 4.2CHNO31.52CFe(III)0.62CFeS20.59e−53723/RTt
Thus, the processes of sulfide minerals dissolution under investigated conditions are limited by internal diffusion [53]. The assessment was based on the obtained Ea values, orders of the reactions with respect to the reactants, SCM equations fitting and SEM-EDS investigation of the samples. Pyrite was proved as an effective catalytic surface for the reduction of nitrate ions and iron (III) with empirical order less than 1.

6. Conclusions

The current work was undertaken to deepen the understanding of the nature of the dissolution process for sulfides Cu12As4S13, CuFeS2 and ZnS in HNO3 media with application FeS2 and Fe (III) ions as catalysts.
It was observed that HNO3 concentration and temperature have the most significant influence on the conversion degree of Cu12As4S13, CuFeS2 and ZnS. The values of Ea were calculated as 28.8, 33.7 and 53.7 kJ/mol, respectively for Cu12As4S13, CuFeS2 and ZnS.
SEM-EDS scanning of the solid residues showed a presence of S° layer covering the surface of the minerals. The latter combined with Ea values and orders of the reactions with respect to the reactants obtained as well as SCM equations fitting allowed us to propose that the dissolution processes are of a diffusion nature.
It was additionally demonstrated that the presence of FeS2 in the system accelerates the conversion process due to galvanic coupling between minerals.
The results obtained can be used in predicting hydrometallurgical processes for sulfide materials such as copper–arsenic ores and concentrates treatment in HNO3 media.
Further detailed kinetic studies on the dissolution of sulfide minerals in HNO3 media such as Cu3AsS4, Cu12Sb4S13, Sb2S3, Cu5FeS4 are of great interest. Furthermore, the complex processing of the low-grade sulfide raw materials in HNO3 media is associated with the extraction of arsenic into the solution, which necessitates the following neutralization of nitrous gases as well as the arsenic utilization in the form of environmentally friendly compounds. These studies are of high relevance in terms of creating industrial hydrometallurgical technology.

Author Contributions

Conceptualization, D.R. and K.K.; methodology, O.D.; validation, O.D. and K.K.; formal analysis, O.D. and A.K.; investigation, O.D., K.K. and D.R.; resources, D.R.; data curation, K.K. and D.R.; writing—original draft preparation, O.D. and K.K.; writing—review and editing, O.D. and A.K.; visualization, O.D. and A.K.; supervision, K.K.; project administration, D.R.; funding acquisition, D.R. All authors have read and agreed to the published version of the manuscript.

Funding

The research funding from the Ministry of Science and Higher Education of the Russian Federation (Ural Federal University Program of Development within the Priority-2030 Program) is gratefully acknowledged.

Acknowledgments

Technicians at Ural Branch of the Russian Academy of Sciences are acknowledged for their assistance with SEM, EDX, analysis.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

Abbreviation/SymbolDescription
1EaActivation energy
2SCMShrinking core model
3XRDX-ray diffraction
4SEM-EDS (or EDX)Scanning electron microscopy/energy dispersive X-ray spectrometry
5EDS-mapping (or EDX-mapping)X-ray characterization technique that allows extremely rapid elemental concentrations to be gathered and collected as map
6ICP-MSInductively coupled plasma mass spectrometry
7XFraction reacted (decimal share)
8EConversion (%)
9Synt. concSynthetic concentrate
10FeS2/synt. concMass ratio of pyrite to synthetic concentrate in a sample of material before the start of the experiment run
11HNO3Nitric acid
12Cu12As4S13Tennantite
13CuFeS2Chalcopyrite
14ZnSSphalerite
15FeS2Pyrite
16FeAsSArsenopyrite
17Elemental sulphur

References

  1. Alguacil, F. The Chemistry of Gold Extraction, 2nd ed.; Marsden, J.O., House, C.I., Eds.; Gold Bulletin; SME: Littleton, CO, USA, 2006; Volume 39, p. 138. [Google Scholar] [CrossRef] [Green Version]
  2. Vikentev, A. Selenium, tellurium and precious metal mineralogy in Uchalinsk copper-zinc-pyritic district, the Urals. IOP Conf. Ser. Mater. Sci. Eng. 2016, 123, 012027. [Google Scholar] [CrossRef] [Green Version]
  3. Shibayama, A.; Takasaki, Y.; William, T.; Yamatodani, A.; Higuchi, Y.; Sunagawa, S.; Ono, E. Treatment of smelting residue for arsenic removal and recovery of copper using pyro–hydrometallurgical process. J. Hazard. Mater. 2010, 181, 1016–1023. [Google Scholar] [CrossRef]
  4. Jarošíková, A.; Ettler, V.; Mihaljevič, M.; Penížek, V.; Matoušek, T.; Culka, A.; Drahota, P. Transformation of arsenic-rich copper smelter flue dust in contrasting soils: A 2-year field experiment. Environ. Pollut. 2018, 237, 83–92. [Google Scholar] [CrossRef]
  5. Lv, X.-D.; Li, G.; Xin, Y.-T.; Yan, K.; Yi, Y. Selective Leaching of Arsenic from High-Arsenic Dust in the Alkaline System and its Prediction Model Using Artificial Neural Network. Min. Metall. Explor. 2021, 28, 2133–2144. [Google Scholar] [CrossRef]
  6. Guo, X.-Y.; Yi, Y.; Shi, J.; Tian, Q.-H. Leaching behavior of metals from high-arsenic dust by NaOH-Na2S alkaline leaching. Trans. Nonferrous Met. Soc. China 2016, 26, 575–580. [Google Scholar] [CrossRef]
  7. Isabaev, S.; Kuzgibekova, K.; Zikanova, T.; Zhinova, E. Complex hydrometallurgical processing of lead arsenic-containing dust from copper production. Tsvetnye Met. 2017, 8, 33–38. [Google Scholar] [CrossRef]
  8. Mamyachenkov, S.V.; Anisimova, O.S.; Kostina, D.A. Improving the precipitation of arsenic trisulfide from washing waters of sulfuric-acid production of copper smelteries. Russ. J. Non-Ferr. Met. 2017, 58, 212–217. [Google Scholar] [CrossRef]
  9. Montenegro, V.; Sano, H.; Fujisawa, T. Recirculation of high arsenic content copper smelting dust to smelting and converting processes. Miner. Eng. 2013, 49, 184–189. [Google Scholar] [CrossRef]
  10. Montenegro, V.; Sano, H.; Fujisawa, T. Recirculation of chilean copper smelting dust with high arsenic content to the smelting process. Mater. Trans. 2008, 49, 2112–2118. [Google Scholar] [CrossRef]
  11. Xue, J.; Long, D.; Zhong, H.; Wang, S.; Liu, L. Comprehensive recovery of arsenic and antimony from arsenic-rich copper smelter dust. J. Hazard. Mater. 2021, 413, 125365. [Google Scholar] [CrossRef]
  12. Karimov, K.; Naboichenko, S.; Kritskii, A.; Tret’yak, M.; Kovyazin, A. Oxidation Sulfuric Acid Autoclave Leaching of Copper Smelting Production Fine Dust. Metallurgist 2019, 62, 1244–1249. [Google Scholar] [CrossRef]
  13. Riveros, P.; Dutrizac, J. The leaching of tennantite, tetrahedrite and enargite in acidic sulfate and chloride media. Can. Metall. Q. 2008, 47, 235–244. [Google Scholar] [CrossRef]
  14. Correia, M.; Carvalho, J.; Monhemius, J. The Effect of Tetrahedrite Composition on Its Leaching Behaviour in FeCl3-NaCl-HCl Solutions. Miner. Eng. 2001, 14, 185–195. [Google Scholar] [CrossRef]
  15. Lin, H. Electrochemical Behavior of Tennantite in Chloride Solutions. J. Electrochem. Soc. 2006, 153, 74–79. [Google Scholar] [CrossRef]
  16. Nazari, A.M.; Radzinski, R.; Ghahreman, A. Review of Arsenic Metallurgy: Treatment of Arsenical Minerals and the Immobilization of Arsenic. Hydrometallurgy 2017, 174, 258–281. [Google Scholar] [CrossRef]
  17. Aghazadeh, S.; Abdollahi, H.; Gharabaghi, M.; Mirmohammadi, M. Selective leaching of antimony from tetrahedrite rich concentrate using alkaline sulfide solution with experimental design: Optimization and kinetic studies. J. Taiwan Inst. Chem. Eng. 2021, 119, 298–312. [Google Scholar] [CrossRef]
  18. Awe, S.; Sandström, A. Selective leaching of arsenic and antimony from a tetrahedrite rich complex sulphide concentrate using alkaline sulphide solution. Miner. Eng. 2010, 23, 1227–1236. [Google Scholar] [CrossRef]
  19. Awe, S.; Samuelsson, C.; Sandström, A. Dissolution kinetics of tetrahedrite mineral in alkaline sulphide media. Hydrometallurgy 2010, 103, 167–172. [Google Scholar] [CrossRef]
  20. Neiva Correia, M.; Carvalho, J.; Monhemius, A. Study of the autoclave leaching of a tetrahedrite concentrate. Miner. Eng. 1993, 11, 1117–1125. [Google Scholar] [CrossRef]
  21. Kritskii, A.; Naboichenko, S.; Karimov, K.; Agarwal, V.; Lundström, M. Hydrothermal Pretreatment of Chalcopyrite Concentrate with Copper Sulfate Solution. Hydrometallurgy 2020, 195, 105359. [Google Scholar] [CrossRef]
  22. Kritskii, A.; Naboichenko, S. Hydrothermal Treatment of Arsenopyrite Particles with CuSO4 Solution. Materials 2021, 23, 7472. [Google Scholar] [CrossRef] [PubMed]
  23. Corkhill, C.; Wincott, P.; Lloyd, J.; Vaughan, D. The oxidative dissolution of arsenopyrite (FeAsS) and enargite (Cu3AsS4) by Leptospirillum ferrooxidans. Geochim. Cosmochim. Acta 2008, 72, 5616–5633. [Google Scholar] [CrossRef]
  24. Xu, R.; Li, Q.; Meng, F.; Yang, Y.; Xu, B.; Yin, H.; Jiang, T. Bio-Oxidation of a Double Refractory Gold Ore and Investigation of Preg-Robbing of Gold from Thiourea Solution. Metals 2020, 10, 1216. [Google Scholar] [CrossRef]
  25. Kritskii, A.; Naboichenko, S.; Klyshnikov, A.; Musaev, V. Pressure Oxidative Leaching of Chalcopyrite Concentrates: Influence of the Process Temperature on Cakes Cyanidation Efficiency. Tsvetnye Met. 2020, 4, 25–29. [Google Scholar] [CrossRef]
  26. Safarzadeh, M.; Moats, M.; Miller, J. Recent trends in the processing of enargite concentrates. Miner. Process. Extr. Metall. Rev. 2014, 35, 283–367. [Google Scholar] [CrossRef]
  27. Anderson, C.; Twidwell, L. Hydrometallurgical processing of gold-bearing copper enargite concentrates. Can. Metall. Q. 2008, 47, 337–346. [Google Scholar] [CrossRef]
  28. Rogozhnikov, D.; Shoppert, A.; Karimova, L. Investigating of nitric acid leaching of high-sulfur copper concentrate. Solid State Phenom. 2020, 229, 968–973. [Google Scholar] [CrossRef]
  29. Shinoda, K.; Tanno, T.; Fujita, T.; Suzuki, S. Coprecipitation of Large Scorodite Particles from Aqueous Fe(II) and As(V) Solution by Oxygen Injection. Mater. Trans. 2009, 50, 1196–1201. [Google Scholar] [CrossRef] [Green Version]
  30. Karimov, K.; Rogozhnikov, D.; Kuzas, E.; Dizer, O.; Golovkin, D.; Tretiak, M. Deposition of Arsenic from Nitric Acid Leaching Solutions of Gold—Arsenic Sulphide Concentrates. Metals 2021, 11, 889. [Google Scholar] [CrossRef]
  31. Tang, Z.; Tang, X.; Liu, H.; Xiao, Z. A Clean Process for Recovering Antimony from Arsenic-Bearing Crystals and Immobilizing Arsenic as Scorodite. Sep. Purif. Technol. 2022, 295, 121276. [Google Scholar] [CrossRef]
  32. Langmuir, D.; Mahoney, J.; Rowson, J. Solubility products of amorphous ferric arsenate and crystalline scorodite (FeAsO4⋯2H2O) and their application to arsenic behavior in buried mine tailings. Geochim. Cosmochim. Acta 2006, 70, 2942–2956. [Google Scholar] [CrossRef]
  33. Paktunc, D.; Dutrizac, J.; Gertsman, V. Synthesis and phase transformations involving scorodite, ferric arsenate and arsenical ferrihydrite: Implications for arsenic mobility. Geochim. Cosmochim. Acta 2008, 75, 2649–2672. [Google Scholar] [CrossRef]
  34. Dixon, D.; Mayne, D.; Baxter, K. Galvanox™—A novel galvanically-assisted atmospheric leaching technology for copper concentrates. Can. Metall. Q. 2008, 47, 327–336. [Google Scholar] [CrossRef]
  35. Nazari, G.; Dixon, D.; Dreisinger, D. Enhancing the kinetics of chalcopyrite leaching in the GalvanoxTM process. Hydrometallurgy 2011, 105, 251–258. [Google Scholar] [CrossRef]
  36. Nazari, G.; Dixon, D.; Dreisinger, D. The mechanism of chalcopyrite leaching in the presence of silver-enhanced pyrite in the GalvanoxTM process. Hydrometallurgy 2012, 113–114, 122–130. [Google Scholar] [CrossRef]
  37. Nazari, G.; Dixon, D.; Dreisinger, D. The role of silver-enhanced pyrite in enhancing the electrical conductivity of sulfur product layer during chalcopyrite leaching in the GalvanoxTM process. Hydrometallurgy 2012, 113–114, 177–184. [Google Scholar] [CrossRef]
  38. Córdoba, E.; Muñoz, J.; Blázquez, M.; González, F.; Ballester, A. Leaching of chalcopyrite with ferric ion. Part I: General aspects. Hydrometallurgy 2008, 93, 81–87. [Google Scholar] [CrossRef]
  39. Yu, S.; Yang, B.; Fang, C.; Zhang, Y.; Liu, S.; Zhang, Y.; Shen, L.; Xie, J.; Wang, J. Dissolution Mechanism of the Oxidation Process of Covellite by Ferric and Ferrous Ions. Hydrometallurgy 2021, 201, 105585. [Google Scholar] [CrossRef]
  40. Hidalgo, T.; Kuhar, L.; Beinlich, A.; Putnis, A. Kinetics and mineralogical analysis of copper dissolution from a bornite/chalcopyrite composite sample in ferric chloride and methanesulfonic-acid solutions. Hydrometallurgy 2019, 188, 140–156. [Google Scholar] [CrossRef]
  41. Lorenzo-Tallafigo, J.; Iglesias-Gonzalez, N.; Romero, R.; Mazuelos, A.; Carranza, F. Ferric leaching of the sphalerite contained in a bulk concentrate: Kinetic study. Miner. Eng. 2018, 125, 50–59. [Google Scholar] [CrossRef]
  42. Neil, C.; Jun, Y. 2016. Fe3+ addition promotes arsenopyrite dissolution and iron (III) (Hydr)oxide formation and phase transformation. Environ. Sci. Technol. Lett 2016, 3, 30–35. [Google Scholar] [CrossRef]
  43. Lorenzo-Tallafigo, J.; Romero-García, A.; Iglesias-González, N.; Mazuelos, A.; Romero, R.; Carranza, F. A Novel Hydrometallurgical Treatment for the Recovery of Copper, Zinc, Lead and Silver from Bulk Concentrates. Hydrometallurgy 2021, 200, 105548. [Google Scholar] [CrossRef]
  44. Dizer, O.; Rogozhnikov, D.; Karimov, K.; Kuzas, E.; Suntsov, A. Nitric Acid Dissolution of Tennantite, Chalcopyrite and Sphalerite in the Presence of Fe (III) Ions and FeS2. Materials 2022, 15, 1545. [Google Scholar] [CrossRef] [PubMed]
  45. Rogozhnikov, D.; Karimov, K.; Shoppert, A.; Dizer, O.; Naboichenko, S. Kinetics and mechanism of arsenopyrite leaching in nitric acid solutions in the presence of pyrite and Fe(III) ions. Hydrometallurgy 2021, 199, 105525. [Google Scholar] [CrossRef]
  46. Levenspiel, O. Chemical Reaction Engineering. Ind. Eng. Chem. Res. 1999, 38, 4140–4143. [Google Scholar] [CrossRef]
  47. Baba, A.; Adekola, F. Hydrometallurgical Processing of a Nigerian Sphalerite in Hydrochloric Acid: Characterization and Dissolution Kinetics. Hydrometallurgy 2010, 101, 69–75. [Google Scholar] [CrossRef]
  48. Gok, O.; Anderson, C.; Cicekli, G.; Cöcen, I. Leaching Kinetics of Copper from Chalcopyrite Concentrate In Nitrous-Sulfuric Acid. Physicochem. Probl. Miner. Process. 2013, 50, 399. [Google Scholar]
  49. Barlett, R. Upgrading copper concentrate by hydrothermal converting chalcopyrite to digenite. Metall. Trans. B 1992, 23, 241–248. [Google Scholar] [CrossRef]
  50. Fuentes, G.; Viñals, J.; Herreros, O. Hydrothermal purification and enrichment of Chilean copper concentrates Part 1: The behavior of bornite, covellite and pyrite. Hydrometallurgy 2009, 95, 104–112. [Google Scholar] [CrossRef]
  51. Kritskii, A.; Celep, O.; Yazici, E.; Deveci, H.; Naboichenko, S. Hydrothermal treatment of sphalerite and pyrite particles with CuSO4 solution. Miner. Eng. 2022, 180, 107507. [Google Scholar] [CrossRef]
  52. Viñals, J.; Fuentes, G.; Hernández, M. Transformation of sphalerite particles into copper sulfide particles by hydrothermal treatment with Cu(II) ions. Hydrometallurgy 2004, 75, 177–187. [Google Scholar] [CrossRef]
  53. Dickinson, C.; Heal, G. Solid-Liquid Diffusion Controlled Rate Equations. Thermochim. Acta 1999, 340–341, 89–103. [Google Scholar] [CrossRef]
Figure 1. The XRD pattern of the minerals: Cu12As4S13 (A), CuFeS2 (B), ZnS (C) and FeS2 (D).
Figure 1. The XRD pattern of the minerals: Cu12As4S13 (A), CuFeS2 (B), ZnS (C) and FeS2 (D).
Materials 15 08149 g001
Figure 2. Effect of temperature on the conversion of Cu12As4S13 (a), CuFeS2 (b), ZnS (c). (300 rpm; 6 mol/L HNO3; 0.018 mol/L Fe (III); FeS2/synt. conc = 1/1).
Figure 2. Effect of temperature on the conversion of Cu12As4S13 (a), CuFeS2 (b), ZnS (c). (300 rpm; 6 mol/L HNO3; 0.018 mol/L Fe (III); FeS2/synt. conc = 1/1).
Materials 15 08149 g002
Figure 3. Effect of HNO3 concentration on the conversion of Cu12As4S13 (a), CuFeS2 (b) and ZnS (c) (300 rpm; 353 K; 0.018 mol/L Fe (III); FeS2/synt. conc = 1/1).
Figure 3. Effect of HNO3 concentration on the conversion of Cu12As4S13 (a), CuFeS2 (b) and ZnS (c) (300 rpm; 353 K; 0.018 mol/L Fe (III); FeS2/synt. conc = 1/1).
Materials 15 08149 g003
Figure 4. The effect of Fe (III) concentration on the conversion of Cu12As4S13 (a), CuFeS2 (b) and ZnS (c) (300 rpm; 353 K; 6 mol/L HNO3; FeS2/synt. conc = 1/1).
Figure 4. The effect of Fe (III) concentration on the conversion of Cu12As4S13 (a), CuFeS2 (b) and ZnS (c) (300 rpm; 353 K; 6 mol/L HNO3; FeS2/synt. conc = 1/1).
Materials 15 08149 g004
Figure 5. The effect of FeS2 additive on the conversion of Cu12As4S13 (a), CuFeS2 (b) and ZnS (c) (300 rpm; 353 K; 6 mol/L HNO3; 0.018 mol/L Fe (III)).
Figure 5. The effect of FeS2 additive on the conversion of Cu12As4S13 (a), CuFeS2 (b) and ZnS (c) (300 rpm; 353 K; 6 mol/L HNO3; 0.018 mol/L Fe (III)).
Materials 15 08149 g005
Figure 6. SEM images of synthetic concentrate-leaching residue (a,b) and EDX-mapping of the residue for Fe (c), Cu (d), S (e) and combined (f). (300 rpm; 353 K; 6 mol/L H2SO4; 0.018 mol/L Fe (III); 60 min; no FeS2 additive).
Figure 6. SEM images of synthetic concentrate-leaching residue (a,b) and EDX-mapping of the residue for Fe (c), Cu (d), S (e) and combined (f). (300 rpm; 353 K; 6 mol/L H2SO4; 0.018 mol/L Fe (III); 60 min; no FeS2 additive).
Materials 15 08149 g006
Figure 7. SEM images of synthetic concentrate-leaching residue (a,b) and EDX-mapping of the residue for Fe (c), Cu (d), S (e) and combined (f). (300 rpm; 353 K; 6 mol/L H2SO4; 0.018 mol/L Fe (III); 60 min; FeS2/synt. conc = 1/1).
Figure 7. SEM images of synthetic concentrate-leaching residue (a,b) and EDX-mapping of the residue for Fe (c), Cu (d), S (e) and combined (f). (300 rpm; 353 K; 6 mol/L H2SO4; 0.018 mol/L Fe (III); 60 min; FeS2/synt. conc = 1/1).
Materials 15 08149 g007aMaterials 15 08149 g007b
Figure 8. XRD patterns after synthetic concentrate dissolution in HNO3 solution without (A) and with (B) the additive of FeS2. (300 rpm; 353 K; 6 mol/L H2SO4; 0.018 mol/L Fe (III); 60 min).
Figure 8. XRD patterns after synthetic concentrate dissolution in HNO3 solution without (A) and with (B) the additive of FeS2. (300 rpm; 353 K; 6 mol/L H2SO4; 0.018 mol/L Fe (III); 60 min).
Materials 15 08149 g008
Figure 9. SEM images of the solid residue after synthetic concentrate dissolution in HNO3 solution. (300 rpm; 353 K; 6 mol/L H2SO4; 0.018 mol/L Fe (III); 15 min).
Figure 9. SEM images of the solid residue after synthetic concentrate dissolution in HNO3 solution. (300 rpm; 353 K; 6 mol/L H2SO4; 0.018 mol/L Fe (III); 15 min).
Materials 15 08149 g009
Figure 10. The linear relationship between 1 – 3(1 – X)2/3 + 2(1 – X) = k·τ and treatment time of Cu12As4S13 (a), CuFeS2 (b) and ZnS (c) at various temperatures.
Figure 10. The linear relationship between 1 – 3(1 – X)2/3 + 2(1 – X) = k·τ and treatment time of Cu12As4S13 (a), CuFeS2 (b) and ZnS (c) at various temperatures.
Materials 15 08149 g010
Figure 11. Arrhenius plots for Cu12As4S13 (a), CuFeS2 (b) and ZnS (c).
Figure 11. Arrhenius plots for Cu12As4S13 (a), CuFeS2 (b) and ZnS (c).
Materials 15 08149 g011
Figure 12. Relationship between SCM equation and CHNO3·CFe(III)·CFeS2·exp[−Ea/(R·T)]·τ·103 for the treatment of Cu12As4S13 (a), CuFeS2(b) and ZnS (c).
Figure 12. Relationship between SCM equation and CHNO3·CFe(III)·CFeS2·exp[−Ea/(R·T)]·τ·103 for the treatment of Cu12As4S13 (a), CuFeS2(b) and ZnS (c).
Materials 15 08149 g012
Table 1. Chemical composition of the synthetic concentrate (%).
Table 1. Chemical composition of the synthetic concentrate (%).
CuFeSAsZn
41.87.1530.213.27.45
Table 2. Normalized EDX analysis results.
Table 2. Normalized EDX analysis results.
ElementFeCuAsSsulfideS0Total
Point 00140.24.21.154.56.8100.0
Point 00243.82.30.853.12.0100.0
Point 00311.322.14.662.042.7100.0
Point 00423.421.71.653.324.3100.0
Point 0057.614.33.374.861.5100.0
Point 00629.411.92.356.219.1100.0
Point 00744.42.10.952.60.5100.0
Point 00845.31.50.352.90.6100.0
Point 0099.123.18.659.236.8100.0
Point 0109.530.410.449.724.4100.0
Table 3. Typical SCM kinetic equations applied for systems with spherical particles.
Table 3. Typical SCM kinetic equations applied for systems with spherical particles.
Limiting StageFormula
1Diffusion through the product layer (sp)1 − 3(1 − X) 2/3 + 2(1 − X)
2Surface chemical reaction (sp)1 − (1 − X)1/3
Table 4. The reaction orders with respect to HNO3 concentration, Fe (III) ions concentration and the amount of FeS2 additive at Cu12As4S13, CuFeS2 and ZnS treatment.
Table 4. The reaction orders with respect to HNO3 concentration, Fe (III) ions concentration and the amount of FeS2 additive at Cu12As4S13, CuFeS2 and ZnS treatment.
Cu12As4S13CuFeS2ZnS
HNO3 concentration
1.21.41.6
Fe(III) ions concentration
0.340.820.62
Amount of FeS2
0.470.690.59
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Dizer, O.; Karimov, K.; Kritskii, A.; Rogozhnikov, D. Synthetic Sulfide Concentrate Dissolution Kinetics in HNO3 Media. Materials 2022, 15, 8149. https://doi.org/10.3390/ma15228149

AMA Style

Dizer O, Karimov K, Kritskii A, Rogozhnikov D. Synthetic Sulfide Concentrate Dissolution Kinetics in HNO3 Media. Materials. 2022; 15(22):8149. https://doi.org/10.3390/ma15228149

Chicago/Turabian Style

Dizer, Oleg, Kirill Karimov, Aleksei Kritskii, and Denis Rogozhnikov. 2022. "Synthetic Sulfide Concentrate Dissolution Kinetics in HNO3 Media" Materials 15, no. 22: 8149. https://doi.org/10.3390/ma15228149

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