Next Article in Journal
Native Seedling Colonization on Stockpiled Mine Soils Is Constrained by Site Conditions and Competition with Exotic Species
Previous Article in Journal
Thermodynamic and Experimental Study on Efficient Extraction of Valuable Metals from Polymetallic Nodules
Previous Article in Special Issue
The High Temperature Co-Processing of Nickel Sulfide and Nickel Laterite Sources
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Direct Acid Leaching of Sphalerite: An Approach Comparative and Kinetics Analysis

by
Nallely G. Picazo-Rodríguez
1,
Ma. de Jesus Soria-Aguilar
2,
Antonia Martínez-Luévanos
1,
Isaias Almaguer-Guzmán
3,
Josue Chaidez-Félix
3 and
Francisco Raul Carrillo-Pedroza
2,*
1
Facultad de Ciencias Químicas, Universidad Autónoma de Coahuila, República S/N, C.P 25280 Saltillo, Coahuila, Mexico
2
Facultad de Metalurgia, Universidad Autónoma de Coahuila, Carretera 57 Km 5, C.P 25710 Monclova, Coahuila, Mexico
3
CIDT-SEPSA Peñoles, C.P. 27268 Torreon, Coahuila, Mexico
*
Author to whom correspondence should be addressed.
Minerals 2020, 10(4), 359; https://doi.org/10.3390/min10040359
Submission received: 26 February 2020 / Revised: 14 April 2020 / Accepted: 15 April 2020 / Published: 17 April 2020
(This article belongs to the Special Issue Hydrometallurgical Processing of Base Metal Sulphides)

Abstract

:
The present work reports the direct leaching of zinc from a sphalerite concentrate in acid media. Lab-scale and pilot-scale experiments were conducted in atmospheric-pressure and low-pressure reactors, respectively. Leaching of zinc and precipitation of iron was achieved in the same stage using different reagents like Fe3+, O2, O3, and Fe2+ (which is continuously oxidized in the leaching solution by H2O2 and O2). The highest percentage of zinc extraction (96%) was obtained in pilot-scale experiments using H2SO4, Fe2+, and O2. Experimental results were compared with those of other researchers to provide a better understanding of the factors influencing the dissolution of zinc. In the first instance, it was determined from analysis of variance that leaching time and the use of an oxidant agent (O2 or O3) were the most influential factors during the direct leaching of zinc from the sphalerite concentrate. Kinetic models were also evaluated to determine the rate-limiting step of the sphalerite leaching; it was concluded that the type of the sulfur layer formed in the residue (porous or non-porous) depends on the type of the oxidant used in the leaching media, which determines the dissolution kinetics of zinc.

1. Introduction

Amongst the non-ferrous metals, zinc is one of the most used metals after aluminum and copper. Zinc is an important metal required for many applications, from metal products to chemical, paint, and agriculture industries. It is mainly recovered from sphalerite (ZnS), which is found in metamorphic, igneous, and sedimentary rocks in many parts of the world [1,2].
The conventional method to recover zinc from sphalerite concentrates involves roasting, leaching, precipitation, purification, and electrolysis. The roasting stage transforms metal sulfides into calcines suitable for hydrometallurgical treatment [3,4,5]. Precipitation of iron is necessary prior to zinc recovery since it causes electrochemical short circuits affecting the electrolytic deposition of zinc [6,7,8].
Although the conventional method for the recovery of zinc has been used over several years, direct leaching of sphalerite concentrates has recently aroused great interest [9,10,11,12,13,14,15,16,17,18,19]. The direct leaching process omits the roasting stage; in this process, elemental sulfur is produced instead of sulfur dioxide, and effective zinc extraction can be achieved independently of the iron concentration in the zinc concentrates [5].
In general, direct leaching processes can be carried out in pressure leaching autoclaves or in agitated tanks operating at atmospheric pressure [20,21]. Atmospheric direct leaching is less expensive than the direct pressure leaching process due to lower maintenance costs [11]. Thereby, there is an increased interest in the use of atmospheric pressure leaching [22]. However, this later is characterized by slow dissolution kinetics. To overcome this shortcoming, researchers have used different oxidants such as ferric ions [12], ozone [22], hydrogen peroxide [23], ammonium, and sodium and potassium persulfate [24].
On the other hand, heterogeneous reactions in which a fluid (gas or liquid) contacts a solid, reacts with it and transforms it into the product, may essentially be represented by the following equation [25]:
A(fluid) +B(solid) → product
Acid leaching follows Equation (1) and can be described by the unreacted shrinking core model [23], where the fluid first reacts with the reactant material at the surface, and then the reaction zone moves inward of the reactant. The unreacted solid sequentially shrinks, while the solid product of the reaction remains or is generally around the same particle and is called ash. The size of the unreacted core decreases as the reaction progresses. In this model, different phases or stages may occur in the reaction with the rate-determining step the one slowest rate in the reaction mechanism [25].
Several kinetic studies of direct leaching of sphalerite, particularly in atmospheric conditions, have demonstrated a two-stage process: A fast leaching rate in the beginning followed by a decrease of the reaction rate after 1 h. The kinetics in the first stage is affected by the chemical reaction and the diffusion rate through the sulfur layer. In this step, the contribution of the chemical reaction decreases gradually as the temperature increases. In the second stage, the rate of sphalerite leaching is controlled only by diffusion through the product layer [26,27,28].
The effect of the leaching solution composition on the dissolution of zinc from sphalerite ore has been investigated at atmospheric pressure using a ferric sulfate acid solution [12]. It was found that the formed sulfur layer inhibited the reaction due to the formation of a denser layer of elemental sulfur on the particle surface. Other researchers studied the direct leaching of sphalerite at atmospheric pressure and acid media using ozone as oxidant; high extractions of zinc (~100%) were obtained accompanied by the formation of a porous sulfur layer [22], which differs from the sulfur layer morphology reported by other researchers [12].
In the present investigation, zinc extraction experiments from a sphalerite concentrate with precipitation of iron using the direct leaching process at lab-scale and pilot-scale were conducted. A parametric study was carried out also using the results obtained by other researchers under different leaching conditions to evaluate the factors influencing the leaching process of zinc concentrates. A kinetic study was also conducted to determine the model and the rate-determining step from the experimental results and other data reported in the literature. It is important to mention that the present work focuses only on the zinc leaching step.

2. Materials and Methods

2.1. Lab-Scale Experiments

Lab-scale experiments of direct leaching of zinc from sphalerite concentrate (D80 particle size = 100 µm) were carried out in a 0.5 M sulfuric acid solution at 353 K under atmospheric pressure. Deionized, distilled water was used in all experiments. Heating was achieved by using a hotplate/stirrer (Thermolyne Cimarec 2, Street Woburn, MA, USA). A 2 L reactor (glass vessel) (Pyrexglass vessel; Corning Inc., New York, NY, USA) was equipped with baffles to prevent swirling of the fluid and fitted with an acrylic sheet cover to minimize loss through evaporation from the leaching solution. The stirring rate remained constant at 500 rpm using a mechanical stirrer type propeller. In some experiments, industrial-grade oxygen (99.5%) or ozone (3% w/w), using an ozone generator (Pacific Ozone L23, Pacific Ozone, Benicia, CA, USA), was injected into the leach liquor through a bubbler placed at the bottom of the reactor. Solid residues were characterized by inductively coupled plasma optical emission spectrometry (Perkin Elmer 8300, Perkin Elmer, Waltham, MA, USA; LECO SC230DR, LECO Corporation, St. Joseph, MI, USA), and X-ray diffraction analysis was performed using a Brucker D8 diffractometer with Cu X-ray tube (Bruker, Billerica, MA, USA). Table 1 shows the conditions of experiments conducted at a laboratory scale.

2.2. Pilot-Scale Experiments

Pilot-scale experiments of direct leaching of zinc from sphalerite concentrate (D80 particle size = 100 µm) were conducted in a 200 L, 316 L stainless steel reactor, which can operate at 0.15 MPa. It was equipped with a chemical seal, a stirring system and 4 deflectors, a controlled cooling/heating jacket, a pressure transmitter and resistance temperature detector, and a Graphtec Data logger.
For the experiments, deionized, distilled water was used; boiling water was added to the reactor to reach a temperature of 368 K, followed by a low speed stirring and the addition of all the materials needed for leaching. A 2-min air purge was made after adding the materials, and industrial grade oxygen (99.5%) was used to increase the pressure of the reactor to 0.096 MPa. It is worth mentioning that in these tests, NaOH was added in such an amount that iron precipitated as jarosite. Such precipitation occurs, as seen later, once the acid has decreased (from pH 2) by its consumption in the leaching reaction of sphalerite. In this work only the leaching stage was considered. Finally, the stirring speed was set at 500 rpm, before starting the data logger to record data over the leaching time.
To study the kinetics of the leaching process of sphalerite, 100 mL samples were taken out from the reactor at different times. Liquids was analyzed by inductively coupled plasma optical emission spectrometry (Perkin Elmer 8300 and LECO SC230DR). Solids were characterized by X-ray fluorescence spectrophotometry (Rigaku Primus II, Rigaku, Tokyo, Japan), X-ray diffraction Brucker D8 diffractometer (Brucker) and scanning electron microscopy (FEI Quanta 600, Thermo Fisher Scientific, Waltham, MA, USA). Table 2 shows the conditions of experiments conducted at a pilot-scale.
The chemical composition of the sphalerite concentrates (used for both the lab-scale and the pilot-scale experiments) is shown in Table 3.
Table 4 shows the main species of the zinc concentrates. The mineralogical composition was realized using the HSC 8 (HSC Chemistry Software) Species Converter module [29] with the data from the chemical analysis by X-ray fluorescence spectrophotometry and the phases determined by X-ray diffraction.

3. Results

3.1. Statistical Analysis

In order to study the effects of the experimental conditions on the extraction of zinc from direct leaching of sphalerite, the results obtained in the present work were compared with the results of other researchers. Obviously, mineralogy plays a very important role in hydrometallurgical processes. In this case, direct leaching studies were identified where the concentration of sphalerite was similar, and with no or low presence of marmatite. This meant that the focus was the effect on the leaching parameters and their impact on the formation and characteristics of the sulfur layer. In turn, it was possible to compare the kinetics of the reaction. Table 5 and Table 6 show the chemical analysis and mineralogical composition of the sphalerite concentrates used in previous studies that were compared with the present work.
Analysis of variance (ANOVA) was conducted to evaluate the individual effect of selected leaching factors on the extraction of zinc. Table 7 shows the factors used for statistical analysis. Factors and levels used for ANOVA were: Leaching time (2 and 5 h), reactor type (1 = opened, 2 = closed), [H2SO4] (0.5, 1.0, 1.5 and 2 M), [Fe2+] (0, 0.05, 0.2 and 0.25 M), [Fe3+] (0, 0.25, 0.5 and 1.2 M), O2/O3 (1 = without oxygen or ozone as oxidants, 2 = oxygen and 3 = ozone), temperature (301, 353, 363 and 365 K) and pulp density (50, 100, 166 and 200 g/L).
Table 8 presents the main effects as a function of zinc extraction, as determined by ANOVA. According to the F-value and the levels of significance (p-values), it can be stated that under the conditions studied, leaching time was the most influential factor followed by the use of O2 or O3, pulp density, and Fe3+ concentration, which had a similar significant effect.
Figure 1 shows the individual effects of the most significant factors on the extraction of Zn: Leaching time, non-use or use of O2 or O3, pulp density, and concentration of Fe3+. As can be seen, the efficiency of Zn extraction increased from 52% to 65% when the leaching time was increased from 2 to 5 h (Figure 1a). Increasing the pulp density from 50 to 166 g/L resulted in an increase of the leaching process efficiency from 60% to 85%, but it decreased up to 30% for a pulp density of 200 g/L. This result was consistent with the results of other researchers [16], who reported incomplete dissolution of metals present in the concentrate during autoclave batch-type experiments carried out at atmospheric pressure with high pulp densities. It was suggested that the increase in the percentage of extraction at a pulp density of 166 g/L was primarily determined by the change in temperature and type of reactor used. However, at sufficiently high pulp density, diffusion of reactants appears to be limited given the increasing concentration of products of the reaction.
With respect to the oxidation of sphalerite, several authors reported the overall chemical reaction of the leaching process [10,11,12,13,14,15,16,17,18,19] as follows (Gibbs energy values were obtained from HSC Chemistry 6 [29]:
ZnS + H2SO4+ 1/2O2(aq) → ZnSO4(aq)+ S0+ H2O; ΔG°80 °C = −251 kJ
ZnS + 2H+ + 0.5O2(aq) → Zn2+ + S0 + H2O; ΔG°80 °C = −186 kJ
Some authors [20] reported that leaching of sphalerite occurred mainly according to Equation (2). Other researchers [18] reported that the sphalerite dissolution rate in acid media was controlled by oxygen diffusion through a porous surface layer of polysulfide or S0. In addition, this equation has been represented, as indicated by Equation (3). We can also consider the chemical equilibrium of the species (Figure 2, from Medusa Software) [31].
According to these figures, the most stable products of sulfuric acid dissociation were H+ and HSO4. In addition, under batch-test conditions (Figure 2a) or equimolar (pilot test) conditions (Figure 2b), the primary product formed from ZnS leaching was ZnSO4 aqueous. Considering these predictions, the following equation is proposed:
ZnS + H+ + HSO4 + 0.5O2(aq) → ZnSO4(aq) + S0 + H2O; ΔG°80 °C= −188 kJ
Therefore, based upon the Gibbs energy, values of Equation (2) and the information given in Figure 2, the leaching of sphalerite was indicated to occur by zinc sulfation and sulfur oxidation. This indicated the involvement of the sulfate ion (or HSO4 anion), which, along with the ZnSO4 ion (both larger molecules), must diffuse through the sulfur layer. Therefore, diffusion of these species can be a limiting step of the reaction.
On the other hand, the chemical theory generally accepted for leaching of ZnS can be described by a process in which ferric sulfate is reduced to ferrous sulfate and elemental sulfur and sphalerite dissolves forming zinc sulfate (aqueous) according to Equation (5) [32,33]. Equation (6) shows only the metallic ion indicating a Gibbs free energy value lower than Equation (5). This indicates that the sulfur is oxidized by the ferric sulfate species, and that the sulfate ion allows the formation of zinc sulfate as aqueous species.
ZnS + Fe2(SO4)3(aq) → ZnSO4(aq)+ 2FeSO4(aq) + S0; ΔG°80 °C= −127 kJ
ZnS + 2Fe3+ → Zn2+ + 2Fe2+ + S0; ΔG°80 °C = −109 kJ
Ferrous iron oxidizes to ferric iron in the presence of oxygen, according to Equation (7). However, comparing the energies values between Equations (7) and (8), its shows that the oxidation reaction is more thermodynamically possible in the presence of sulphuric acid rather than just H+.
2FeSO4(a)+ H2SO4 + 1/2O2(aq) → Fe2(SO4)3(aq)+ H2O; ΔG°80 °C = −124 kJ
2Fe2++ 2H+ + 1/2O2(aq) → 2Fe3+ + H2O; ΔG°80 °C = −77 kJ
According to previous equations, ferric iron reacts with ZnS and is reduced to ferrous ion, which the oxidizes to ferric ion when it reacts with oxygen present in the leaching system; these reaction favor oxidant conditions during leaching [11]. The thermodynamics seems to indicate that the formation of zinc sulfate (aqueous) is more favored than the zinc ion.
As observed in Figure 1c, low zinc extractions (50%) were obtained in the direct leaching of sphalerite without added Fe3+. However, the efficiency increased up to 75% with additions of Fe3+; this percentage was independent of the amount of Fe3+ concentration (0.25 M to 1.2 M). Figure 1d shows the variations in the zinc extractions as a function of the oxidant (1 = without O2 or O3 as oxidants, 2 = O2, and 3 = O3) present in the leaching media. The reactions between the sphalerite and O2 is indicated in Equation (2), while the reaction of sphalerite with O3 occurs according to the following:
ZnS + H2SO4 + 1/3O3(aq) → ZnSO4(aq) + S0 + H2O; ΔG°80 °C = −296.84 kJ
ZnS + 2H+ + 1/3O3(aq) → Zn2+ + S0 + H2O; ΔG°80 °C = −237.48 kJ
Again, formation of zinc sulfate aqueous (Equation (9)) resulted in a higher Gibbs energy value than the formation of zinc as an ion (Equation (10)). As can be seen, experiments without any gas or with O2 lead to an average zinc extraction of 65%. On the other hand, the use of O3 (the strongest oxidant) caused a decrease in zinc extraction to 45%. Figure 1d shows the average of the extraction results of experiments 3 and 4 (Table 7) and Mubarok et al. [22] tests. It is notable that the latter [22] used ozone without Fe3+ at room temperature, while experiments 3 and 4 were run at 80 °C. The low average zinc extraction may be due to the averaging of data obtained from the use of ozone under high temperature leaching conditions (where the stability and solubility of ozone is lower). The conclusion of Mubarok et al. [22] was that ozone promoted fast oxidation of Fe2+ to Fe3+ and that constant availability of Fe3+ may promote the formation of a non-porous sulfur layer and, therefore, inhibit the extraction of zinc as the reaction proceeds. This and the lower stability of ozone at a higher temperature may explain the low zinc extractions obtained in this study.

3.2. Kinetic analysis

The kinetics of the leaching process was determined from zinc extraction data as a function of time obtained from lab-scale experiments (experiment 1) and pilot-scale experiments (experiments 5, 6, and 7), and it was described by the shrinking core model.
Table 9 shows a compilation of experimental data obtained in the present work and data reported in the literature made to determine the rate-determining step of the dissolution of zinc under different conditions of leaching.
The rate-limiting step of the zinc dissolution was obtained from a comparison between experimental data and some kinetic models reported in the literature [34], as shown in Table 10. Figure 3 shows the leaching fraction (X) as a function of time for experiments shown in Table 9.
The kinetics analysis was conducted to compare the models reported in the literature using the zinc extraction data (reacted fraction) from Figure 3. Figure 4 shows a comparison of the reaction rate of experiments shown in Table 7, which is represented by the slope corresponding to each curve. It is worth mentioning that this figure was obtained after plotting the corresponding model for each test. The data obtained from each model were normalized in the range of 0 to 1 to be able to visualize and compare the slope of each line. Normalization consisted simply of adjusting the values obtained from each model; for example, if the model gave values on the order of 10−2, it was multiplied by 10, thus that the values were in the order of 10−1. In the case of 10−3, by 100 for the same effect, and so on. Being multiples, the slope of each model was conserved, which allowed a comparison of values of the same order (from 10−1) on the Y-axis. Each slope line was proportional to the apparent constant of the reaction rate of the experiments studied. The higher the slope, the higher the reaction rate. As can be observed, the reaction rate of experiments P1, P2, P5 and P6 was slower than P3, P4, P7, and P8.
All the experiments presented in Table 9 were adjusted to the shrinking model, however, differences were found in the controlling stages. Experiment P1 fits well to the mechanism that involves the product layer model, according to the following equation.
Kt = 1 − (1 − X)2/3
The kinetics of the leaching process in experiments P5 and P6 (conducted at laboratory scale and at pilot-scale, respectively), is described by the following model:
Kt = 1 − (1 − X)2/3
Some authors have reported that the sphalerite reacts mainly with Fe3+ ions forming a surface product layer, which is probably dense [12,22].
In the case of these experiments, Equations (11) and (12) that represent the oxidation cycle of iron, can be rapid at the beginning, which causes the rapid formation of the elemental sulfur layer due to the type of solid-liquid reaction. As can be seen in Table 7, both tests were conducted in a batch reactor without supplying additional oxygen to the leaching system, therefore, this inhibits the speed of Equations (3) and (4), and consequently, the rate of dissolution of zinc.
Results of experiments P3 and P4 carried out at pilot-scale showed that the chemical reaction was the rate-determining step of the direct leaching of sphalerite concentrates. It suggested that the product layer formed at the surface of mineral particles was porous, thin or, that it can be easily detached from the particle and not limiting for the diffusion of reactants and reaction products. The fraction leached as a function of time in these experiments, which exhibited chemically controlled kinetic behavior is expressed as follows:
Kt = 1 − (1 − X)1/3
The kinetics of the leaching process in experiments P5 and P6 (conducted at laboratory scale and at pilot-scale, respectively), is described by the following model:
Kt = (1/(1 − X)1/3 − 1)2
This model relates the diffusion process where the concentration of the reactants decreases as the reaction progresses. This can occur when there is a relatively high percentage of solids within the solution.
The use of a high pulp density (200 g/L) in experiment P5 leads to a slow dissolution rate, which may be associated with the low availability of reactants with respect to solids in the leaching media. Furthermore, in experiment P5, Fe2+ ions were used, without supplying oxygen to the leaching medium, an initial amount of H2O2 was added, which could have caused the rapid formation of a dense sulfur layer on the surface of the sphalerite particles, in a similar way to experiments P1 and P2.
Experiment P6 conducted in a low-pressure reactor in the presence of sulfuric acid, and oxygen exhibited the presence of H2S. Initially, oxygen consumption was observed; then it stopped, and negative pressure was observed (a characteristic that a gas is leaving the same oxygen inlet line). When purging the line, the H2S gas came out (detected by its smell). Some researchers reported that the dissolution of sphalerite follows the indirect oxidation reactions involving acid dissolution (Equation (15)), the release of H2S, and its further oxidation to elemental sulfur by oxygen or ferric ions (Equation (16)) present in the leaching medium. Therefore, in this case, the chemical reaction in the interface is supposed to be the rate-determining step of the sphalerite leaching process [19].
ZnS + H2SO4 → ZnSO4(aq)+H2S (aq); △G°95 °C = −37.96 kJ
H2S (aq) + 1/2O2 → S0 +H2O; △G°95 °C = −200.05 kJ
H2S(aq) + Fe2(SO4)3(aq) → H2SO4 +2FeSO4(aq)+ S0; △G°95 °C = −93.964 kJ
H2S(aq) + 2Fe3+ → 2H++2Fe2+ + S0; △G°95 °C = −138.1 kJ
It is possible that the formation of H2S in experiment P6 was faster than its oxidation, which may have caused its accumulation in the reactor, making the passage of oxygen difficult and inhibiting the rate of zinc dissolution.
The leaching medium used during experiments P3 and P4 may be the cause of the formation of a porous sulfur layer, since Fe2+ and O2 ions were used. Based on the results of the previous experiments, it was inferred that during the P3 and P4 experiments, the reaction of the Fe3+ ions with the sphalerite occurred in a controlled manner, since it involved first the oxidation reaction of Fe2+ to Fe3+ and in secondly the reaction of Fe3+ with the sphalerite. Oxygen (supplied into the pulp) reacts with sphalerite and in the oxidation-reduction cycle of iron, which can be the cause of the pores observed in the sulfur layer, which allows a greater dissolution of zinc. SEM characterization of the leaching residues (P3 and P4 experiments) showed a porous layer of sulfur on the surface of the unreacted sphalerite particles, as shown in Figure 5 and Figure 6.
Cháidez et al. [35] carried out a direct acid leaching of chalcopyrite in the same type of reactor used in experiments P3 and P4 and observed a type of porous sulfur layer on unreacted chalcopyrite particles.
Table 11 shows the mineralogical species present in the residue of experiment P4. As can be observed, the percentage of sulfur increases from 3% at t = 0 h up to 56% at t = 4 h. In addition, despite the formation of a layer of elemental sulfur (apparently porous), 4 h later, most of the zinc has already been dissolved.
The chemical reaction is also the rate-determining step of experiments P7 and P8 [22]. In these cases, elemental sulfur present in the leaching residue floated easily, forming independent particles, which were released on the mineral surface instead of promoting the formation of layers on the reacted particle. Unlike the atmospheric leaching of sphalerite concentrates with ferric sulfates, there was no apparent obstacle for the dissolution of zinc during the entire leaching process [5,12]. Mubarok et al. [22] also reported that the chemical reaction was the rate-determining step of the sphalerite leaching reaction, which was consistent with the kinetic analysis conducted in the present work.

4. Conclusions

Acid leaching of metal sulfides has been extensively studied, mainly chalcopyrite, pyrite, and sphalerite, among other minerals. The consensus in the literature is that the dissolution reaction of the polymetallic sulfides has in common the formation of elemental sulfur. This product is formed around the surface of the mineral. The formed layer may be porous, non-porous, or, in some cases, detached from the surface. Therefore, the elemental sulfur layer determines the controlling step of the dissolution reaction of the metal.
In this context, the present study evaluated some of the variables used for sphalerite leaching. The data and results of laboratory and pilot-scale experiments obtained in this work were used and compared with the literature. The analysis of variance (ANOVA) suggests that the use of highly oxidizing conditions can promote the formation of a non-porous layer of sulfur.
In the kinetic study, it was found that when Fe3+ is used as the initial oxidant, the model that limits zinc dissolution is the diffusion of reactants and products, because this oxidant promotes the formation of a layer of non-porous sulfur.
The best result obtained in this work was 96% zinc extraction after 5 h, which was obtained to pilot-scale in a pressure low-pressure reactor under the following conditions: 353 K, 0.096 MPa of pressure, 350 rpm, 1.5 M H2SO4, 200 g/L of pulp density and 0.2 M Fe2+.
The high extraction of zinc obtained in this investigation is related to the reagents used in this case, Fe2+ oxidized with O2 to obtain Fe3+, where the speed of this oxidation reaction is slower than the dissolution of zinc, therefore, this chemical reaction limits the reaction rate. This indicates that the elemental sulfur layer formed is porous, as was demonstrated in the micrographs presented in the results section.

Author Contributions

Conceptualization, I.A.-G. and F.R.C.-P.; methodology, F.R.C.-P., I.A.-G. and N.G.P.-R.; software, N.G.P.-R.; validation, I.A.-G., J.C.-F. and M.d.J.S.-A.; formal analysis, N.G.P.-R., M.d.J.S.-A., F.R.C.-P. and A.M.-L.; investigation, N.G.P.-R.; resources, I.A.-G. and F.R.C.-P.; data curation, M.d.J.S.-A., J.C.-F. and A.M.-L.; writing—original draft preparation, N.G.P.-R.; writing—review and editing, F.R.C.-P.; visualization, F.R.C.-P., I.A.-G., J.C.-F., M.d.J.S.-A. and N.G.P.-R.; supervision, I.A.-G.; project administration, F.R.C.-P. and I.A.-G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

Research and Development Center (CIDT-SEPSA) Peñoles supplied the sphalerite concentrate samples, chemical analysis, DRX, and SEM characterization. N. Picazo appreciates the CONACYT scholarship for their PhD thesis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dutrizac, J.E.; Macdonald, J.C. The dissolution of sphalerite in ferric chloride solution. Metall. Trans. B 1978, 9B, 543–551. [Google Scholar] [CrossRef]
  2. Habashi, F. The Metal Industry Ferrous Metals; WILEY-VCH: Weinheim, Germany, 1997; Volume I. [Google Scholar]
  3. Bobeck, G.E.; Su, H. The kinetics of dissolution of sphalerite in ferric chloride solution. Metall. Trans. B 1985, 16B, 413–424. [Google Scholar] [CrossRef]
  4. Sahu, S.K.; Sahu, K.K.; Pandey, B.D. Leaching of zinc sulfide concentrate from the ganesh–himal deposit of Nepal. Metall. Mater. Trans. B 2006, 37, 541–549. [Google Scholar] [CrossRef]
  5. Souza, A.D.; Pina, P.S.; Leão, V.A. Bioleaching and chemical leaching as an integrated process in the zinc industry. Miner. Eng. 2007, 20, 591–599. [Google Scholar] [CrossRef] [Green Version]
  6. Ismael, M.R.; Carvalho, J.M. Iron recovery from sulphate leach liquors in zinc hydrometallurgy. Miner. Eng. 2003, 16, 31–39. [Google Scholar] [CrossRef]
  7. Sayilgan, E.; Kukrer, T.; Civelekoglu, G.; Ferella, F.; Akcil, A.; Veglio, F.; Kitis, M. A review of technologies for the recovery of metals from spent alkaline and zinc–carbon batteries. Hydrometallurgy 2009, 97, 158–166. [Google Scholar] [CrossRef]
  8. Tsakiridis, P.E.; Oustadakis, P.; Katsiapi, A.; Agatzini, S. Hydrometallurgical process for zinc recovery from electric arc furnace dust (EAFD). Part II: Downstream processing and zinc recovery by electrowinning. J. Hazard. Mater 2010, 179, 8–14. [Google Scholar] [CrossRef]
  9. Dutrizac, J.E. The dissolution of sphalerite in ferric sulfate media. Metall. Mater. Trans. B 2005, 37, 161–171. [Google Scholar] [CrossRef]
  10. Karimi, S.; Rashchi, F.; Moghaddam, J. Parameters optimization and kinetics of direct atmospheric leaching of Angouran sphalerite. Int. J. Miner. Process. 2017, 162, 58–68. [Google Scholar] [CrossRef]
  11. Lampinen, M.; Laari, A.; Ilkka, T. Kinetic model for direct leaching of zinc sulfide concentrates at high slurry and solute concentration. Hydrometallurgy 2015, 153, 160–169. [Google Scholar] [CrossRef]
  12. Lochmann, J.; Pedĺik, M. Kinetic anomalies of dissolution of sphalerite in ferric sulfate solution. Hydrometallurgy 1995, 37, 89–96. [Google Scholar] [CrossRef]
  13. Lorenzo, J.; Iglesias, N.; Romero, R.; Mazuelos, A. Ferric leaching of the sphalerite contained in a bulk concentrate: Kinetic study. Miner. Eng. 2018, 125, 50–59. [Google Scholar] [CrossRef]
  14. Peng, P.; Xie, H.; Lua, L. Coupling leaching of sphalerite concentrate. Miner. Eng. 2005, 18, 553–555. [Google Scholar] [CrossRef]
  15. Santos, S.M.; Ismael, M.R.; Correia, M.J.; Reis, M.T.; Deep, A.; Carvalho, J.M. Hydrometallurgical treatment of a zinc concentrate by atmospheric direct leach process. In Proceedings of the European Congress of Chemical Engineering, Copenhagen, Denmark, 16–20 September 2007; pp. 16–20. [Google Scholar]
  16. Santos, S.M.; Machado, R.M.; Correia, M.J.; Reis, M.T.; Ismael, M.R.; Carvalho, J.M. Ferric sulphate/chloride of zinc and minor elements from a sphalerite concentrate. Miner. Eng. 2010, 23, 606–615. [Google Scholar] [CrossRef]
  17. Svens, K. Direct leaching alternatives for zinc concentrates. In Chen Honorary Symposium on Hydrometallurgy, Electrometallurgy and Aterials Characterization, 1st ed.; Shijie, W., Dutrizac, J.E., Free, M.L., Hwang, J.Y., Kim, D., Eds.; Wiley: Hoboken, NJ, USA, 2012; pp. 191–206. [Google Scholar] [CrossRef]
  18. Weisener, C.G.; Smart, R.S.; Gerson, A.R. A comparison of the kinetics and mechanism of acid leaching of sphalerite conintaining low and high concentrations of iron. Int. J. Miner. Process. 2004, 74, 239–249. [Google Scholar] [CrossRef]
  19. Xu, Z.; Jiang, Q.; Wang, C. Atmospheric oxygen-rich direct leaching behavior of zinc sulphide concentrate. Trans. Nonferrous Met. Soc. China 2013, 23, 3780–3787. [Google Scholar] [CrossRef]
  20. Buban, K.R.; Collins, M.J.; Masters, I.M.; Trytten, L.C. Comparison of direct pressure leaching with atmospheric leaching of zinc concentrates. In Lead-Zinc, 1st ed.; Dutrizac, J.E., Gonzalez, J.A., Henke, D.M., James, S.E., Siegmund, A.H., Eds.; Wiley: Hoboken, NJ, USA, 2000; pp. 727–738. [Google Scholar] [CrossRef]
  21. Lampinen, M.; Laari, A.; Turunen, I. Simulation of direct leaching of zinc concentrate in a CSTR. Can. J. Chem. Eng. 2010, 88, 625–632. [Google Scholar] [CrossRef]
  22. Mubarok, M.Z.; Sukamoto, K.; Ichlas, Z.T.; Sugiarto, A.T. Direct sulfuric acid leaching of zinc sulfide concentrate using ozone as oxidant under atmospheric pressure. Min. Met. Process. 2018, 35, 133–140. [Google Scholar] [CrossRef]
  23. Pecina, T.; Franco, T.; Castillo, P.; Orrantia, E. Leaching of a zinc concentrate in H2SO4 solutions containing H2O2 and complexing agents. Miner. Eng. 2008, 21, 23–30. [Google Scholar] [CrossRef]
  24. Babu, M.N.; Sahu, K.K.; Pandey, B.D. Zinc recovery from sphalerite concentrate by direct oxidative leaching with ammonium, sodium and potassium persulphates. Hydrometallurgy 2002, 64, 119–129. [Google Scholar] [CrossRef]
  25. Levenspiel, O. Chemical reaction engineering, 3rd ed.; Wiley: New York, NY, USA, 1999; ISBN 9780471254249. [Google Scholar]
  26. Acero, P.; Cama, J.; Ayora, C. Sphalerite dissolution kinetics in acidic environment. Hydrometallurgy 2007, 22, 1872–1883. [Google Scholar] [CrossRef]
  27. Da Silva, G. Relative importance of diffusion and reaction control during the bacterial and ferric sulphate leaching of zinc sulphide. Hydrometallurgy 2004, 73, 313–324. [Google Scholar] [CrossRef]
  28. Markus, H. Reduction of ferric to ferrous with sphalerite concentrate, kinetic modelling. Hydrometallurgy 2004, 73, 269–282. [Google Scholar] [CrossRef]
  29. Roine, A. HSC Chemistry 8. Available online: www.outotec.com/hsc (accessed on 22 March 2019).
  30. Pedroza, S. Estudio de la Reducción y Precipitación de Hierro de la Solución Ácida de la Lixiviación del Zinc. Ph.D. Thesis, Universidad Autómoma de Coahuila, Saltillo, Mexico, 2018. [Google Scholar]
  31. Puigdomenech, I. Medusa Software. Available online: www.kth.se/che/medusa (accessed on 6 April 2020).
  32. Corriou, J.P.; Gély, R.; Viers, P. Thermodynamic and kinetic study of the pressure leaching of zinc sulfide in aqueous sulfuric acid. Hydrometallurgy 1988, 21, 85–102. [Google Scholar] [CrossRef]
  33. Dehghan, R.; Noaparast, M.; Kolahdoozan, M.; Mousavi, S.M. Statistical evaluation and optimization of factors affecting the leaching performance of a sphalerite concentrate. Int. J. Miner. Process. 2008, 89, 9–16. [Google Scholar] [CrossRef]
  34. Grénman, H.; Salmi, T.; Murzin, D.Y. Solid-liquid reaction kinetics-experimental aspects and model development. Rev. Chem. Eng. 2011, 27, 53–77. [Google Scholar] [CrossRef]
  35. Cháidez, J.; Parga, J.; Valenzuela, J.; Carrillo, R.; Almaguer, I. Leaching Chalcopyrite Concentrate with Oxygen and Sulfuric Acid Using a Low-Pressure Reactor. Metals 2019, 9, 189. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Effect of leaching parameters on the efficiency of zinc extraction: (a) Leaching time, (b) pulp density, (c) Fe3+ concentration, (d) oxidant (1 = without O2 or O3, 2 = O2 and 3 = O3).
Figure 1. Effect of leaching parameters on the efficiency of zinc extraction: (a) Leaching time, (b) pulp density, (c) Fe3+ concentration, (d) oxidant (1 = without O2 or O3, 2 = O2 and 3 = O3).
Minerals 10 00359 g001
Figure 2. Chemical equilibrium diagrams using Medusa software [31]: (a) Batch-test conditions and (b) pilot test conditions.
Figure 2. Chemical equilibrium diagrams using Medusa software [31]: (a) Batch-test conditions and (b) pilot test conditions.
Minerals 10 00359 g002
Figure 3. Reacted fraction as a function of time for experiments shown in Table 7.
Figure 3. Reacted fraction as a function of time for experiments shown in Table 7.
Minerals 10 00359 g003
Figure 4. Normalized values of each kinetic model shown in Table 9.
Figure 4. Normalized values of each kinetic model shown in Table 9.
Minerals 10 00359 g004
Figure 5. SEM images of leaching residues obtained from experiment P3 showing unreacted sphalerite particles surrounded by elemental sulfur.
Figure 5. SEM images of leaching residues obtained from experiment P3 showing unreacted sphalerite particles surrounded by elemental sulfur.
Minerals 10 00359 g005
Figure 6. SEM images of leaching residues obtained from experiment P4 showing the presence of pyrite, sulfur, and sphalerite particles surrounded by natrojarosite.
Figure 6. SEM images of leaching residues obtained from experiment P4 showing the presence of pyrite, sulfur, and sphalerite particles surrounded by natrojarosite.
Minerals 10 00359 g006
Table 1. Experimental conditions used for the lab-scale experiments of direct leaching of zinc from sphalerite concentrate.
Table 1. Experimental conditions used for the lab-scale experiments of direct leaching of zinc from sphalerite concentrate.
Oxidant[Fe2+], MPulp Density, g/L
H2O2 (50%)0.05200
O2 (99.5%)0.0550
O3 (3%)050
O3 (3%) 0.0550
Table 2. Experimental conditions used at pilot-scale experiments of direct leaching of zinc from sphalerite concentrate.
Table 2. Experimental conditions used at pilot-scale experiments of direct leaching of zinc from sphalerite concentrate.
TimeTemperatureAgitation RatepO2,Pulp Density,
hKrpmMPaMg/LMM
123685000.0961.52000.20.2
103685000.0961.51660.040.25
53685000.0961.614000
Table 3. Chemical analysis of sphalerite concentrates (wt.%).
Table 3. Chemical analysis of sphalerite concentrates (wt.%).
Ag (g/t)CaCdCuFePbSSiZn
41.180.340.380.6311.70.5223.11.251.5
Table 4. Mineralogical composition of the sphalerite concentrates (wt.%).
Table 4. Mineralogical composition of the sphalerite concentrates (wt.%).
Identified SpeciesComposition (wt.%)
Sphalerite(Zn, Fe) S85
ChalcopyriteCuFeS22.0
PyriteFeS21.1
GalenaPbS0.5
QuartzSiO22.4
Others 9
Table 5. Chemical analysis of the sphalerite concentrates (reported in the literature) and that of the sample used for this work.
Table 5. Chemical analysis of the sphalerite concentrates (reported in the literature) and that of the sample used for this work.
SampleChemical Analysis (wt.%)
ZnFeCuPbS
Souza et al. [5]51.49.0 1.931.8
Santos et al. [16]42.89.92.11.5
Karimi et al. [10]49.112.240.133.54
Lampinen et al. [11]54.26.6 1.530.3
Mubarok et al. [22]51.773.094.551.832.57
Table 6. Mineralogical composition of the sphalerite concentrate (reported in the literature) and that of the sample used for this work.
Table 6. Mineralogical composition of the sphalerite concentrate (reported in the literature) and that of the sample used for this work.
SampleSphalerite, ZnsPyrite, FeSChalcopyrite, CuFeS₂Galena, PbSQuartz, SiO₂
Souza et al. [5] 95****
Santos et al. [16]82.210.83.6**
Karimi et al. [10]73****
Lampinen et al. [11]*2*1.8*
Mubarok et al. [22]88*9.932.07*
Pedroza [30] and this work85*20.53
* Information not available or reported.
Table 7. Experimental conditions of direct leaching of zinc from sphalerite considered for ANOVA (analysis of variance).
Table 7. Experimental conditions of direct leaching of zinc from sphalerite considered for ANOVA (analysis of variance).
ReferenceType of ReactorPressure (MPa)Type of Agitator[H2SO4]M[Fe2+]M[Fe3+]MO2/O3Particle SizeTemperature (K)Pulp Density (g/L)Zn Extraction after 2 h (%)Zn Extraction after 5 h (%)
Souza et al [5]10.0212 propellers (paddle)100.51100% < 44 μm363507090
Santos et al [16]10.0211 propeller (paddle)0.500.5199% < 70 μm353509095
Santos et al [16]10.0211 propeller (paddle)0.500.5199% < 70 μm3531005870
Santos et al [16]10.0211 propeller (paddle)0.500.25299% < 70 μm353507080
Karimi et al [10]10.096Magnetic0.501.21D80 = 23.5 μm353506280
Lampinen [11]20.492 propellers (turbine)1.50.0502D80 = 27 μm3681005085
Mubarok et al [22]10.021Magnetic2.0003100% < 74 μm301505080
Pedroza [30] 10.0961 propeller (helix)0.500.52D80 = 100 μm353505045
* Experiment 1 10.0961 propeller (helix0.50.501D80 = 100 μm3532001616
* Experiment 2 10.0961 propeller (helix)0.50.502D80 = 100 μm353504450
* Experiment 3 10.0961 propeller (helix)0.5003D80 = 100 μm353502945
* Experiment 410.0961 propeller (helix)0.50.503D80 = 100 μm353502134
* Experiment 520.0212 propellers (paddle and hélix)1.50.202D80 = 100 μm3682003259
* Experiment 620.0962 propellers (paddle and hélix)1.50.2502D80 = 100 μm3681667396
* Experiments conducted in the present work.
Table 8. Analysis of variance of data shown in Table 7.
Table 8. Analysis of variance of data shown in Table 7.
Source of VarianceDegrees of FreedomSum SquaresMean SquaresF-Valuep-Value
Leaching time, h115361536.5914.165 × 10−3*
Type of reactor14.89 × 10−114.89 × 10−1100.99
Pulp density, g/L31750583.345.3825 × 10−3*
Temperature, K38.41× 10−122.80 × 10−1201
O2/O3224151207.6711.134 × 10−3*
[H2SO4], M3821273.912.520.13
[Fe2+], M321270.750.650.60
[Fe3+], M31574524.844.480.03*
* Significance level of alpha = 0.05.
Table 9. Experiments of direct leaching of zinc with different initial conditions.
Table 9. Experiments of direct leaching of zinc with different initial conditions.
ExperimentReference[Fe3+] M[Fe2+] M[H2SO4] MPulp Density (g/L)O2/O3Particle SizeZn Extraction at 5 h, %
P1Pedroza [29]0.0600.560NoneD80 = 100 µm47
P2Pedroza [29]0.0500.360NoneD80 = 100 µm28
P3Experiment 5*00.031.4200O2D80 = 100 µm59
P4Experiment 6*00.251.5165.6O2D80 = 100 µm96
P5Experiment 1*00.050.5200NoneD80 = 100 µm16
P6Experiment 7*001.6140O2D80 = 100 µm9
P7Mubarak et al. [22]00250O3100% < 74 µm80
P8Mubarak et al [22]002150O3100% < 74 µm55
* Experiments conducted in the present work.
Table 10. Adjustments of kinetic models with experimental data shown in Table 9.
Table 10. Adjustments of kinetic models with experimental data shown in Table 9.
ExperimentKinetic ModelDescriptionkR2
P11 − (2/3)X − (1 − X)2/3Diffusion through a product layer0.020.92
P21 − (1 − X)2/3Diffusion through a thin film0.060.99
P31 − (1 − X)1/3Chemical reaction0.060.96
P41 − (1 − X)1/3Chemical reaction0.190.96
P5(1/(1 − X)1/3 −1)2Concentration of diffusing species varies with X0.0010.95
P6(1/(1 − X)1/3 − 1)2Concentration of diffusing species varies with X0.00020.97
P71 − (1 − X)1/3Chemical reaction0.080.98
P81 − (1 − X)1/3Chemical reaction0.040.98
Table 11. Mineralogical reconstruction of leaching residues of experiment P4 obtained at different times.
Table 11. Mineralogical reconstruction of leaching residues of experiment P4 obtained at different times.
PhasesComposition (wt.%)
t = 0 ht = 0.1 ht = 0.25 ht = 0.5 h t = 1 ht = 1.5 ht = 2 ht = 3 ht = 4 h
Sphalerite, (Zn, Fe)S857770665847342112
Chalcopyrite, CuFeS2232
Galena, PbS0.5
Pyrite, FeS2 23344444
Anglesite, PbSO4 11111212
Gypsum, CaSO4(H2O)2 68561169
Elemental sulfur, S 35111929354656
Other minerals12.51413111313142217

Share and Cite

MDPI and ACS Style

Picazo-Rodríguez, N.G.; Soria-Aguilar, M.d.J.; Martínez-Luévanos, A.; Almaguer-Guzmán, I.; Chaidez-Félix, J.; Carrillo-Pedroza, F.R. Direct Acid Leaching of Sphalerite: An Approach Comparative and Kinetics Analysis. Minerals 2020, 10, 359. https://doi.org/10.3390/min10040359

AMA Style

Picazo-Rodríguez NG, Soria-Aguilar MdJ, Martínez-Luévanos A, Almaguer-Guzmán I, Chaidez-Félix J, Carrillo-Pedroza FR. Direct Acid Leaching of Sphalerite: An Approach Comparative and Kinetics Analysis. Minerals. 2020; 10(4):359. https://doi.org/10.3390/min10040359

Chicago/Turabian Style

Picazo-Rodríguez, Nallely G., Ma. de Jesus Soria-Aguilar, Antonia Martínez-Luévanos, Isaias Almaguer-Guzmán, Josue Chaidez-Félix, and Francisco Raul Carrillo-Pedroza. 2020. "Direct Acid Leaching of Sphalerite: An Approach Comparative and Kinetics Analysis" Minerals 10, no. 4: 359. https://doi.org/10.3390/min10040359

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