Next Article in Journal
Ore Geology, Fluid Inclusion Microthermometry and H-O-S Isotopes of the Liyuan Gold Deposit, Central Taihang Mountains, North China Craton
Previous Article in Journal
Indium Mineralization in the Volcanic Dome-Hosted Ánimas–Chocaya–Siete Suyos Polymetallic Deposit, Potosí, Bolivia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Quartz on the Preparation of Sodium Stannate from Cassiterite Concentrates by Soda Roasting Process

School of Minerals Processing and Bioengineering, Central South University, Changsha 410083, China
*
Author to whom correspondence should be addressed.
Minerals 2019, 9(10), 605; https://doi.org/10.3390/min9100605
Submission received: 19 August 2019 / Revised: 23 September 2019 / Accepted: 25 September 2019 / Published: 1 October 2019
(This article belongs to the Section Mineral Processing and Extractive Metallurgy)

Abstract

:
Sodium stannate (Na2SnO3) has been successfully prepared by a novel process of roasting cassiterite concentrates and sodium carbonate (Na2CO3) under CO–CO2 atmosphere, namely soda roasting-leaching process. However, more than 22 wt. % tin of the cassiterite was not converted into Na2SnO3 and entered the leach residues. Quartz (SiO2) is the predominant gangue in the cassiterite, and phase evolution of SnO2–SiO2–Na2CO3 system roasted under CO–CO2 atmosphere was still uncertain. In this study, the effect of SiO2 in cassiterite concentrates on preparation of Na2SnO3 was clarified. The results indicated that Na8SnSi6O18 was inevitably formed when cassiterite and Na2CO3 were roasted above 775 °C under CO–CO2 atmosphere via the reaction of SnO2 + 6SiO2 + 4Na2CO3 = Na8SnSi6O18 + 4CO2, and formation of Na8SnSi6O18 would be increased with increasing roasting temperature and Si/Sn mole fraction. In addition, it was found that Na8SnSi6O18 was insoluble in the leachate at pH value range of 1–14, which, therefore, was enriched in the leach residues. The silicon content of the cassiterite concentrates should be controlled as lower as possible to obtain a higher conversion ratio of Na2SnO3.

1. Introduction

Sodium stannate (Na2SnO3) is highly desirable in many fields, including electroplating [1], tin alloy production [2], and solid superbase catalysts for dehydrogenation and flame retardants [3,4]. Recently, it is also used as solid electrolytes and electrode materials in chemical sources of electrical energy [5]. The consumption of sodium stannate has been increasing rapidly in the last decade. The traditional Na2SnO3 preparation processes applied metallic tin and low-melting-point sodium hydroxide (NaOH) as raw materials, which were conducted in a fused state in the presence of sodium nitrate (NaNO3) as oxidizers [6]. However, metallic tin was always obtained from high-temperature reduction smelting process. Besides, some secondary tin-containing resources, including stanniferous alloy, tin scrap, waste solder, and electronic waste, have also been used for preparing sodium stannate [7,8,9,10,11,12,13], and these processes would cause high production cost and long process flow. In addition, the emission of hazardous gases (NH3 and NOx) deriving from the oxidizers (NaNO3) was also a shortcoming.
The authors’ group has developed a novel process for preparing Na2SnO3 from cassiterite concentrates (SnO2) and sodium carbonate (Na2CO3) roasted in a solid-state under CO–CO2 atmosphere [14,15], namely soda roasting–leaching process. Sodium stannate trihydrate (Na2SnO3·3H2O) with a purity of 95.8 wt. % was obtained, which met the requirement of industrial first-grade products. This process displays a bright prospect in preparing Na2SnO3, and the function mechanism and the formation kinetics of Na2SnO3 have also been clarified in previous publications [16,17,18,19]. It was found that CO–CO2 atmosphere promoted the formation of oxygen deficiency on the surface of cassiterite, which broke the stable structure of SnO2. Then the activation energy of the reactions between Na2CO3 and SnO2 decreased significantly, and the formation of Na2SnO3 was much easily under CO–CO2 atmosphere [15,16,17,18,19,20]. Nevertheless, it was found that over 22 wt. % tin of the cassiterite was not converted into Na2SnO3 and disposed as residues during the roasting–leaching process [20].
Quartz (SiO2) is the predominant gangue mineral in cassiterite ores, which cannot be separated perfectly by beneficiation combined methods, such as gravity concentration and froth flotation [21,22]. Before used as raw materials, the cassiterite concentrates should be firstly pretreated by oxidation roasting–acid leaching process to remove the impurity elements, including Fe, As, S, Pb, Sb, etc. [23,24]. However, the quartz (SiO2) is very hard to be removed during the pretreatment process.
Our previous studies have discussed in detail about the formation mechanism of Na2SnO3; however, the reaction principle and thermodynamic data of Na2CO3–SnO2–SiO2 system were unclear [25,26,27]. Hence, the major objectives of this study were: (1) to determine the effect of SiO2 on the leaching efficiency of Sn and Si; (2) to investigate the effect of SiO2 on phase evolution of SnO2–Na2CO3 system; (3) to ascertain leaching characteristics of the tin, silicon-containing compounds.

2. Materials and Methods

2.1. Materials

As described in our previous studies [14], firstly, cassiterite concentrates were roasted in air at 900 °C for 120 min and then leached with 25% HCl to remove the main impurity elements, including Fe, As, S, Pb, and Sb. The chemical compositions of the original and pretreated cassiterite concentrates are given in Table 1. All the testing samples were pre-ground to a particle size passing through a 200 mesh screen (<0.074 mm). The gases used in this study included CO, CO2, and N2 gases, all of which were with purity of 99.99 vol. %.
Analytically pure Na2CO3, SnO2, and SiO2 were also used to investigate the reaction mechanism.

2.2. Methods

2.2.1. Roasting Process

Sodium salt roasting was one of the effective methods to treat minerals [28,29,30,31]. The roasting tests were performed in the high-temperature zone of a horizontal electric resistance furnace. An experimental schematic diagram for the roasting tests is the same as that reported in the previous study [16,32]. The materials were weighed precisely at a certain molar ratio and mixed up gently with an agate mortar and pestle for 30 min. After that, mixed samples were dried in a drying oven at 105 °C for 4 h and then a dried sample about 5.0 g was placed in a corundum crucible (80 mm × 10 mm) and loaded into a heat resistant corundum tube (diameter 45 mm). The crucible carrying the sample was pushed toward the constant roasting zone located in the central area of an electrically heated horizontal tube furnace. Beforehand, N2 gas was introduced into the corundum tube until the temperature reached a constant value. Next, the N2 was immediately replaced by the mixed CO–CO2 gas. Inlet gas flow rate was fixed at 4.0 L/min. The sample was then roasted in a 15 vol. % CO atmosphere at a given temperature (775 °C, 825° C, 875 °C, and 925 °C) for different roasting time (5 min, 10 min, 15 min, 20 min, 30 min, and 60 min). After that, the roasted samples were cooled in pure N2 atmosphere. Finally, the cooled samples were ready for further analyses.

2.2.2. Leaching Process

The leaching tests were conducted in 250 mL round bottom flasks with a mechanical stirring paddle, and the stirring rate was fixed at 300 rpm for each test. A water bath was used to control the temperature at 40 °C. Next, the ground roasted products of 10.0 g and distilled water of 40 mL were put into the flasks and leached at a fixed pH solution for 60 min in the water bath (pH was fixed at 12.5 based on our previous study [14]). Finally, the leaching solution was filtered and prepared for the determination of Sn and Si concentration. The residues were washed with distilled water to identify the phase constituents.
The leaching efficiency of Sn and Si, which is calculated according to the following equation:
L = 1000 C V M W 100 %
where L is the leaching efficiency of Sn or Si, M is the weight of the roasted samples (g), W is the Sn or Si grade of the roasted samples (%), C is the mass concentration of Sn or Si in the leaching solution (mg/mL), and V is the volume of leaching solution (mL).

2.2.3. Instrument Techniques

The chemical compositions of cassiterite concentrates were examined using an X-ray fluorescence spectrometer (XRF, Axios MAX, PANalytical, Almelo, The Netherlands). The phase constituents of the samples were identified by X-ray diffraction (XRD; D/max 2550PC, Rigaku Co. Ltd, Tokyo, Japan) with the step of 0.02° at 10°∙min−1 ranging from 10° to 80°. The content of Sn and Si in the solid material and the aqueous solution were determined using inductively coupled plasma atomic emission spectroscopy (ICP-AES; Icap7400 Radial, Thermo Fisher Scientific, Waltham, MA, USA).

3. Results and Discussion

3.1. Behaviour of Si during the Soda Roasting-Leaching Processs

3.1.1. Effect of Roasting Temperatures and Time

Our previous studies showed that trihydrate sodium stannate (Na2SnO3·3H2O) was obtained by roasting cassiterite concentrates and Na2CO3 in CO–CO2 atmosphere [14], and other impurity elements were almost removed in the pretreatment process. However, the pretreated cassiterite concentrates still contained 3.66 wt. % Si. To examine the impact of SiO2 on the formation of Na2SnO3, the leaching efficiency of Sn and Si was firstly investigated at different roasting temperatures and time.
Figure 1 and Figure 2 illustrate the effect of roasting temperatures and roasting time on leaching efficiency of Sn and Si under the setting experimental conditions: CO content of 15% and Na2CO3/SnO2 mole ratio of 1.5.
It was observed from Figure 1 that the roasting temperatures had significant impact on leaching efficiency of Sn and Si. The leaching efficiency of Sn and Si increased with increasing the roasting temperatures. The leaching efficiency of Sn and Si increased significantly as roasting temperatures increased from 800 °C to 1000 °C. At 1000 °C, the leaching efficiency of Sn and Si was almost the same level, 97.5 wt. % and 96.1 wt. %, respectively. The melting point of Na2CO3 is reported as 851 °C, and melted Na2CO3 is inclined to volatilize, which resulted in equipment corrosion at higher temperatures [33,34]. Therefore, 875 °C is selected to be a suitable roasting temperature based on the previous study [14].
Figure 2 shows the effect of roasting time in the range of 30–120 min on the leaching efficiency of Sn and Si. The leaching efficiency of Si increased from 26.3 wt. % to 70.2 wt. % as the roasting time increased from 30 to 60 min, while the Sn leaching efficiency was higher than 65 wt. % as the roasting time prolonged. The results indicated that reaction rate of Equation (2) was much faster than that of Equation (3) in CO–CO2 atmosphere.
SnO2 + Na2CO3 = Na2SnO3 + CO2
SiO2 + Na2CO3 = Na2SiO3 + CO2
When the roasting time was further extended to 90 min, the leaching efficiency of Sn and Si almost stayed unchanged. As the roasting time was 120 min, the leaching efficiency of Sn and Si was 81.4 wt. % and 76.6 wt. %, respectively.
Based on the results given in Figure 1 and Figure 2, the distribution of Sn and Si elements in the whole experimental flowsheet is presented in Figure 3. The content of Sn and Si in pretreated cassiterite concentrates were 62.9 wt. % and 3.7 wt. %, respectively. Then, the roasted samples were leached in the setting experimental conditions. 70.2 wt. % of Si entered the leachate. Meanwhile, it was worthy to note that leaching efficiency of Sn was 77.7 wt. % and 22.3 wt. % Sn entered the residues. To discover the underlying reason, the phase components of roasted samples and leach residues should be further determined.

3.1.2. Phase Analysis of Roasted Samples and Leach Residues

The XRD pattern raw material (in Figure 3) is shown in Figure 4, which contained tin oxide (SnO2) and silicon dioxide (SiO2). The XRD pattern of Sample 1# (in Figure 3) is shown in Figure 5. The main phases of the roasted samples were Na2SnO3 and Na2SiO3, and the content of Sn and Si was 43.0 wt. % and 3.1 wt. %, respectively (as shown in Figure 3). In particular, the characteristic peaks of Na8SnSi6O18 were also observed in Figure 5. It was noteworthy that the phase of Na8SnSi6O18 was never reported in previous researches.
Sample 1# was then used for the leaching test. The XRD pattern of the leach residues, Sample 2# (in Figure 4), is shown in Figure 6. The content of Sn and Si in the residues was 43.1 wt. % and 2.9 wt. %, as shown in Figure 4. The main phases of the residues were SnO2 and Na8SnSi6O18. Moreover, the diffraction intensities of Na8SnSi6O18 were more intensive compared with those shown in Figure 6, indicating that Na8SnSi6O18 was enriched in the residues. Meanwhile, the diffraction peaks of Na2SnO3 and Na2SiO3 disappeared. Therefore, we drew the following conclusions: (1) a few of SnO2 was were not converted into soluble stannate; (2) Na8SnSi6O18 was formed in the roasting process; (3) Na2SnO3 and Na2SiO3 were almost dissolved into the leachate during the leaching process; (4) Na8SnSi6O18 was possibly insoluble and enriched in the residues.

3.1.3. Solution Chemistry of Metasilicic Acid and Tin

To further determine the existing forms of Sn and Si ions in the leachate, the solution chemistry of leachate with different pH value was analyzed. Dissolution thermodynamics of Sn and Si ions in aqueous solution was firstly discussed in the leaching process. In previous publications [35,36,37], it was reported that Sn(IV) might exist as Sn4+, SnOH3+, Sn(OH)22+, Sn(OH)3+, Sn(OH)4, Sn(OH)5, and Sn(OH)62−, while Si(IV) could exist as H2SiO3, HSiO3, and SiO32−. The mole fractions of metasilicic and stannum species at varying pH values were calculated based on previous studies (reaction constant (K) shown in Table 2) [38,39,40,41], and the results are shown in Figure 7.
Figure 7 illustrated the very strong hydrolysis behaviors of Si(IV) and Sn(IV). The main species of metasilicic and stannum were Sn4+, SnOH3+, Sn(OH)22+, and H2SiO3 at pH value below 2. The neutral Sn(OH)4 and H2SiO3 were dominant at pH value of 2–7.5. At pH of 9.5, the negatively charged HSiO3, SiO32−, Sn(OH)5, and Sn(OH)62− were the main aqueous species. However, after pH value above 12.5, the mole fraction of HSiO3 was lower compared to that of SiO32−. Moreover, Sn(OH)4 and H2SiO3 disappeared. Generally, a leaching solvent with a pH greater than 12 is preferred for the prevention of stannate hydrolysis [14]. Hence, Sn(OH)62− and SiO32− only existed in the solution when the value of pH was more than 12.6 (0.05 mol/L NaOH). Figure 6 also indicated that there were no H2SiO3(s) and Sn(OH)4(s) in the leach residues, illustrating that H2SiO3(s) and Sn(OH)4(s) was not formed during the leaching process.
The possible reaction of Na2SnO3(s) during the leaching process was expressed as Equation (12), and Sn(OH)62− anions were instantly formed because of their high stability under weakly alkaline condition [42]. Na2SiO3(s) was dissolved as forms of Na+ and SiO32− (aq.) by Equation (13) at pH of 12.6 [36]. However, Sn(OH)62− did not react with SiO32− to form Na8SnSi6O18 in the solution with pH value of 12.6. Therefore, we inferred that Na8SnSi6O18 was only formed in the roasting process.
Na2SnO3(s) + 3H2O = Na2Sn(OH)6 (aq.) = 2Na+ + Sn(OH)62−
Na2SiO3(s) = 2Na+ + SiO3(aq)2−
The abovementioned results indicated that SiO2 in the cassiterite concentrates could react with Na2CO3 and SnO2 to form Na2SnO3, Na2SiO3, and Na8SnSi6O18 during the roasting process. Part of SiO2 reacted with SnO2 and Na2CO3 to form Na8SnSi6O18. Moreover, Na2SnO3 and Na2SiO3 were more easily soluble than Na8SnSi6O18. In order to make sure how Na8SnSi6O18 was formed during the roasting process, effect of SiO2 on phase evolution of SnO2–Na2CO3 system was further researched.

3.2. Effect of SiO2 on Phase Evolution of SnO2–Na2CO3 System

It was reported that SiO2/SnO2 molar ratio in cassiterite concentrates was about 1:4 [14]. Hence, in order to research the effect of SiO2 on phase evolution of SnO2–Na2CO3 system, the analytical grade reagents of Na2CO3 and SiO2 were added into cassiterite concentrates with different SiO2/SnO2 molar ratios. In particular, Na2CO3/(SnO2 + SiO2) mole ratio was fixed as 1.5.

3.2.1. Effect of SiO2/SnO2 Mole Ratio

Firstly, different SiO2 dosage was added into cassiterite concentrates in order to investigate the effect of SiO2/SnO2 mole ratio. The XRD patterns of the samples roasted at 875 °C for 60 min in a 15 vol. % CO atmosphere are shown in Figure 8. SiO2/SnO2 mole ratio varied in the range of 1:4–7:1.
As observed from Figure 8, the main phases of the roasted samples were Na2SnO3, Na2SiO3, and a small amount of Na8SnSi6O18 when the SiO2/SnO2 mole ratio was 1:4. As the SiO2/SnO2 mole ratio increased from 1:4 to 7:1, the diffraction peak intensity of Na8SnSi6O18 increased significantly. Meanwhile, the diffraction peak of Na2SnO3 was gradually weakened and vanished. The Sn/Si theoretical value in Na8SnSi6O18 was 1:6. The results indicated that high Si content in cassiterite concentrates promoted the formation of Na8SnSi6O18, which decreased the conversion of SnO2 to Na2SnO3 during the roasting process.

3.2.2. Effect of Roasting Temperature

The effect of roasting temperature was then performed, and Figure 9 shows the XRD patterns of the samples roasted for 60 min at temperature range of 775–925 °C. SiO2/SnO2 mole ratio was fixed as 7:1.
It was found that the phases of the samples roasted at 775 °C were Na8SnSi6O18, Na2SiO3, and a small amount of Na2CO3. The diffraction peak intensities of Na8SnSi6O18 were enhanced with increasing the roasting temperature, while those of Na2SiO3 and Na2CO3 were weakened. When the temperature reached at 875 °C, Na8SnSi6O18 was the predominant substance in the roasted samples. The diffraction peaks of Na2SiO3 and Na2CO3 almost vanished as the roasting temperature increased to 925 °C, and Na8SnSi6O18 was the only phase in the roasted samples.
It was worthy to note that 875 °C was the suitable roasting temperature for Na2SnO3 preparation by the soda roasting–leaching process [14]. Besides, as observed from Figure 4, the leaching efficiency of Sn and Si at 875 °C was 77.7 wt. % and 70.2 wt. %, respectively. Furthermore, the leaching efficiency of Sn and Si was derived from the formation of soluble Na2SnO3 and Na2SiO3. The results further illustrated that a part of SnO2 and SiO2 was transformed into Na8SnSi6O18. The corresponding reaction was expressed as the reaction of Equation (14).
SnO2 + 6SiO2 + 4Na2CO3 = Na8SnSi6O18 + 4CO2

3.3. Leaching Behavior of Na8SnSi6O18

Na8SnSi6O18 was inevitably formed in the roasting process, and it was enriched in the leach residues. However, the leaching behavior of Na8SnSi6O18 was never conducted. In this section, to determine the leaching behavior of Na8SnSi6O18, Na8SnSi6O18 was synthesized under the following conditions: mole ratio of Na2CO3:SnO2:SiO2 = 4:1:6, roasting temperature of 1000 °C, and roasting time of 120 min. The XRD pattern of synthetic Na8SnSi6O18 is shown in Figure 10. The results indicated that the roasted product had a very high purity and was well-matched with the PDF standard card (PDF#72-2449) of Na8SnSi6O18 phase, and there were no diffraction peaks of other impurities.
Effect of pH on leaching behavior of Na8SnSi6O18 was performed under the following leaching experimental conditions: liquid-to-solid ratio of 4 cm3/g, leaching temperature of 40 °C, leaching time of 60 min, and stirring rate of 300 rpm [14].
The effect of pH value on the leaching efficiency of Si and Sn was investigated systematically, and the results are shown in Figure 11a. The leaching efficiency of Sn decreased from 65.1 wt. % to 1.9 wt. % and Si from 21.2 wt. % to 4.9 wt. % as the pH value increased from −0.6 (4 mol/L HCl) to 1. The leaching efficiency of Si and Sn was substantially unchanged as the pH value increased further. The leaching efficiency of Sn and Si was 0.5 wt. % and 1.3 wt. % (in Figure 11a) when the pH value was 12.6 (0.05 mol/L NaOH). The XRD pattern of the leach residues obtained at pH of 12.6 is also presented in Figure 11b. It is seen from Figure 11b that Na8SnSi6O18 was the only phase in the residues. The results verified that Na8SnSi6O18 phase was almost insoluble and residual in the residues.
In the whole, Na8SnSi6O18 was inevitably formed in the roasting process for preparation of Na2SnO3 if using cassiterite as raw materials. Moreover, Na8SnSi6O18 was insoluble in the pH range of 1–14, which was the main reason that part of tin oxides was not converted into soluble Na2SnO3 and enriched in the residues. Thus, if cassiterite concentrates were used as raw materials to prepare Na2SnO3 by the soda roasting–leaching process, the silicon content in the cassiterite should be controlled as low as possible.

4. Conclusions

In this study, the effect of SiO2 on sodium stannate preparation from cassiterite and Na2CO3 roasted under CO–CO2 atmosphere was investigated. The main conclusions were summarized as follows:
  • Na2SnO3, Na2SiO3, and Na8SnSi6O18 were easily formed when the Na2CO3 + SnO2 + SiO2 mixtures were roasted under CO–CO2 atmosphere. Na2SnO3 and Na2SiO3 were easily dissolved into the leachate during the leaching process, while Na8SnSi6O18 enriched into the leach residues.
  • Na8SnSi6O18 was inevitably formed in the roasting process of preparation of Na2SnO3. Roasting temperature and Si/Sn mole ratio were the two critical factors affecting the formation of Na8SnSi6O18, which was more easily formed at higher roasting temperature and Si/Sn mole ratio.
  • The leaching behavior of synthetic Na8SnSi6O18 indicated that Na8SnSi6O18 was almost insoluble in the leachate at the pH range of 1–14. Therefore, the loss of tin in the residues was mainly attributed to the insoluble Na8SnSi6O18 formed during the roasting process.

Author Contributions

Y.Z. conceived the project and wrote the final paper. Z.S. performed the experiments and wrote initial drafts of the work. B.H. performed the experiment. X.C. performed the XRD analysis. M.L., S.L., J.L., and T.J. discussed the content. All authors discussed the results and reviewed the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (No.51574283 and No.51234008).

Acknowledgments

The authors would express their heartful thanks to Financial supports from the National Natural Science Foundation of China and Co-Innovation Center for Clean and Efficient Utilization of Strategic Metal Mineral Resources, and language editing assistance from Corby Anderson in Kroll Institute for Extractive Metallurgy, Colorado School of Mines, USA.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sharma, A.; Das, S.; Das, K. Effect of different electrolytes on the microstructure, corrosion and whisker growth of pulse plated tin coatings. Microelectron. Eng. 2017, 170, 59–68. [Google Scholar] [CrossRef]
  2. Yang, W.; Xu, D.; Yao, X. Stable preparation and characterization of yellow micro arc oxidation coating on magnesium alloy. J. Alloy. Compd. 2018, 745, 609–616. [Google Scholar] [CrossRef]
  3. Zhang, S.G.; Wei, Y.D.; Yin, S.F.; Luo, S.L.; Au, C.T. Superbasic sodium stannate as catalyst for dehydrogenation, Michael addition and transesterification reactions. Appl. Catal. A Gen. 2011, 406, 113–118. [Google Scholar] [CrossRef]
  4. Sang, B.; Li, Z.; Yu, L.; Li, X.; Zhang, Z. Preparation of zinc hydroxystannate-titanate nanotube flame retardant and evaluation its smoke suppression efficiency for flexible polyvinyl chloride matrix. Mater. Lett. 2017, 204, 133–137. [Google Scholar] [CrossRef]
  5. Butova, V.V.; Shukaev, I.L. Ion exchange conversion of solid electrolyte, potassium sodiostannate, into isomorphous metastable sodium stannate. Mendeleev Commun. 2016, 26, 246–247. [Google Scholar] [CrossRef]
  6. Wang, L.; Pu, S. Direct preparation of sodium stannate from tin concentrates. J. Cent. South Inst. Min. Metall. 1987, 18, 427–431. [Google Scholar]
  7. Liu, W.; Li, W.; Han, J.; Wu, D.; Li, Z.; Gu, K.; Qin, W. Preparation of calcium stannate from lead refining slag by alkaline leaching-purification-causticization process. Sep. Purif. Technol. 2019, 212, 119–125. [Google Scholar] [CrossRef]
  8. Wu, D.; Liu, W.; Han, J.; Jiao, F.; Xu, J.; Gu, K.; Qin, W. Direct preparation of sodium stannate from lead refining dross after NaOH roasting-water leaching. Sep. Purif. Technol. 2019, 227, 115683. [Google Scholar] [CrossRef]
  9. Liu, W.; Li, Z.; Han, J.; Li, W.; Wang, X.; Wang, N.; Qin, W. Selective Separation of Arsenic from Lead Smelter Flue Dust by Alkaline Pressure Oxidative Leaching. Minerals 2019, 9, 308. [Google Scholar] [CrossRef]
  10. Wu, D.; Han, J.; Liu, W.; Jiao, F.; Qin, W. Preparation of Calcium Stannate from Lead Refining Dross by Roast–Leach–Precipitation Process. Minerals 2019, 9, 283. [Google Scholar] [CrossRef]
  11. Akimoto, Y.; Iizuka, A.; Shibata, E. High-voltage pulse crushing and physical separation of polycrystalline silicon photovoltaic panels. Miner. Eng. 2018, 125, 1–9. [Google Scholar] [CrossRef]
  12. Yang, C.; Li, J.; Tan, Q.; Liu, L.; Dong, Q. Green process of metal recycling: Coprocessing waste printed circuit boards and spent tin stripping solution. ACS Sustain. Chem. Eng. 2017, 5, 3524–3534. [Google Scholar] [CrossRef]
  13. Yang, C.; Tan, Q.; Liu, L.; Dong, Q.; Li, J. Recycling Tin from electronic waste: A problem that needs more attention. ACS Sustain. Chem. Eng. 2017, 5, 9586–9598. [Google Scholar] [CrossRef]
  14. Zhang, Y.; Su, Z.; Liu, B.; You, Z.; Yang, G.; Li, G.; Jiang, T. Sodium stannate preparation from stannic oxide by a novel soda roasting–leaching process. Hydrometallurgy 2014, 146, 82–88. [Google Scholar] [CrossRef]
  15. Liu, B.; Zhang, Y.; Su, Z.; Li, G.; Jiang, T. Phase Evolution of Tin Oxides Roasted Under CO–CO2 Atmospheres in the Presence of Na2CO3. Miner. Process. Extr. Metall. Rev. 2016, 37, 264–273. [Google Scholar] [CrossRef]
  16. Liu, B.; Zhang, Y.; Su, Z.; Li, G.; Jiang, T. Function mechanism of CO-CO2 atmosphere on the formation of Na2SnO3 from SnO2 and Na2CO3 during the roasting process. Powder Technol. 2016, 301, 102–109. [Google Scholar] [CrossRef]
  17. Su, Z.; Zhang, Y.; Liu, B.; Chen, J.; Li, G.; Jiang, T. Effect of CaCO3 on the Gaseous Reduction of Tin Oxide Under CO-CO2 Atmosphere. Miner. Process. Extr. Metall. Rev. 2016, 37, 179–186. [Google Scholar] [CrossRef]
  18. Su, Z.; Zhang, Y.; Liu, B.; Chen, Y.; Li, G.; Jiang, T. Effect of CaF2 On the Reduction Volatilization of Tin Oxide Under CO-CO2 Atmosphere. Miner. Process. Extr. Metall. Rev. 2017, 38, 207–213. [Google Scholar] [CrossRef]
  19. Liu, B.; Zhang, Y.; Su, Z.; Li, G.; Jiang, T. Formation kinetics of Na2SnO3 from SnO2 and Na2CO3 roasted under CO-CO2 atmosphere. Int. J. Miner. Process. 2017, 165, 34–40. [Google Scholar] [CrossRef]
  20. Zhang, Y.; Su, Z.; You, Z.; Liu, B.; Yang, G.; Li, G.; Jiang, T. Sodium stannate preparation from cassiterite concentrate and sodium carbonate by roasting under a CO–CO2 atmosphere. In Proceedings of the Rare Metal Technology, Conference Proceeding, TMS, San Diego, CA, USA, 16–20 February 2014; pp. 163–169. [Google Scholar]
  21. Feng, Q.; Wen, S.; Zhao, W.; Chen, Y. Effect of calcium ions on adsorption of sodium oleate onto cassiterite and quartz surfaces and implications for their flotation separation. Sep. Purif. Technol. 2018, 200, 300–306. [Google Scholar] [CrossRef]
  22. Tian, M.; Zhang, C.; Han, H.; Liu, R.; Gao, Z.; Chen, P.; He, J.; Hu, Y.; Sun, W.; Yuan, D. Novel insights into adsorption mechanism of benzohydroxamic acid on lead (II)-activated cassiterite surface: An integrated experimental and computational study. Miner. Eng. 2018, 122, 327–338. [Google Scholar] [CrossRef]
  23. Wright, P.A. Extractive Metallurgy of Tin, 2nd ed.; Elsevier: Amsterdam, The Netherlands, 1982. [Google Scholar]
  24. Song, X.C. Tin Metallurgy; Metallurgical Industry Press: Beijing, China, 2011. [Google Scholar]
  25. Paparoni, G.; Webster, J.D.; Walker, D. Experimental techniques for determining tin solubility in silicate melts using silica capsules in 1 atm furnaces and rhenium capsules in thepiston cylinder. Am. Mineral. 2010, 95, 776–783. [Google Scholar] [CrossRef]
  26. Zhang, Y.; Su, Z.; Liu, B.; Zhou, Y.; Jiang, T.; Li, G. Reduction behavior of SnO2 in the tin-bearing iron concentrates under CO-CO2 atmosphere. Part II: Effect of quartz. Powder Technol. 2016, 291, 337–343. [Google Scholar] [CrossRef]
  27. Zhang, Y.; Liu, B.; Su, Z.; Chen, J.; Li, G.; Jiang, T. Effect of Na2CO3 on the preparation of metallic tin from cassiterite roasted under strong reductive atmosphere. J. Min. Metall. 2016, 52, 9–15. [Google Scholar] [CrossRef]
  28. Sanchez-Segado, S.; Monti, T.; Katrib, J.; Kingman, S.; Dodds, C.; Jha, A. Towards sustainable processing of columbite group minerals: Elucidating the relation between dielectric properties and physico-chemical transformations in the mineral phase. Sci. Rep. 2017, 7, 18016. [Google Scholar] [CrossRef]
  29. Escudero-Castejon, L.; Sanchez-Segado, S.; Parirenyatwa, S.; Hara, Y.; Jha, A. A Cr6+-free extraction of chromium oxide from chromite ores using carbothermic reduction in the presence of alkali. In Applications of Process Engineering Principles in Materials Processing, Energy and Environmental Technologies; Springer: Berlin/Heidelberg, Germany, 2017; pp. 179–188. [Google Scholar]
  30. Parirenyatwa, S.; Escudero-Castejon, L.; Sanchez-Segado, S.; Hara, Y.; Jha, A. An investigation on the kinetics and mechanism of alkali reduction of mine waste containing titaniferous minerals for the recovery of metals. In Applications of Process Engineering Principles in Materials Processing, Energy and Environmental Technologies; Springer: Berlin/Heidelberg, Germany, 2017; pp. 465–474. [Google Scholar]
  31. Sanchez-Segado, S.; Lahiri, A.; Jha, A. Alkali roasting of bomar ilmenite: Rare earths recovery and physico-chemical changes. Open Chem. 2015, 13, 270–278. [Google Scholar] [CrossRef]
  32. Su, Z.; Zhang, Y.; Han, B.; Liu, B.; Lu, M.; Peng, Z.; Li, G.; Jiang, T. Synthesis, characterization, and catalytic properties of nano-SnO by chemical vapor transport (CVT) process under CO–CO2 atmosphere. Mater. Des. 2017, 121, 280–287. [Google Scholar] [CrossRef]
  33. Knutsson, P.; Lai, H.; Stiller, K. A method for investigation of hot corrosion by gaseous Na2SO4. Corros. Sci. 2013, 73, 230–236. [Google Scholar] [CrossRef]
  34. Li, Y.; He, J.B.; Zhang, M.; He, X.L. Corrosion inhibition effect of sodium phytate on brass in NaOH media. Potential-resolved formation of soluble corrosion products. Corros. Sci. 2013, 74, 116–122. [Google Scholar] [CrossRef]
  35. Séby, F.; Potin-Gautier, M.; Giffaut, E.; Donard, O.F.X. A critical review of thermodynamic data for inorganic tin species. Geochim. Cosmochim. Acta 2001, 65, 3041–3053. [Google Scholar] [CrossRef]
  36. Park, H.; Englezos, P. Osmotic coefficient data for Na2SiO3 and Na2SiO3-NaOH by an isopiestic method and modeling using Pitzer’s model. Fluid Phase Equilib. 1998, 153, 87–104. [Google Scholar] [CrossRef]
  37. Kim, E.; Osseo-Asare, K. Dissolution windows for hydrometallurgical purification of metallurgical-grade silicon to solar-grade silicon: Eh-pH diagrams for Fe silicides. Hydrometallurgy 2012, 127, 178–186. [Google Scholar] [CrossRef]
  38. Zhao, Z.; Shuai, W.; Zhang, J.; Chen, X. Sn (IV) anions adsorption onto ferric hydroxide: A speciation-based model. Hydrometallurgy 2013, 140, 135–143. [Google Scholar] [CrossRef]
  39. Lindsay, W.L. Chemical Equilibria in Soils; Wiley: New York, NY, USA, 1979; pp. 79–85. [Google Scholar]
  40. Hummel, W.; Berner, U.; Curti, E.; Pearson, F.J.; Thoenen, T. Nagra/PSI chemical thermodynamic data base 01/01. Radiochim. Acta 2002, 90, 805–813. [Google Scholar] [CrossRef]
  41. Angadi, S.I.; Sreenivas, T.; Jeon, H.S.; Baek, S.H.; Mishra, B.K. A review of cassiterite beneficiation fundamentals and plant practices. Miner. Eng. 2015, 70, 178–200. [Google Scholar] [CrossRef]
  42. Choi, Y.I.; Salman, S.; Kuroda, K.; Okido, M. Synergistic corrosion protection for AZ31 Mg alloy by anodizing and stannate post-sealing treatments. Electrochim. Acta 2013, 97, 313–319. [Google Scholar] [CrossRef]
Figure 1. Effect of roasting temperature on leaching efficiency of Sn and Si (Time: 60 min).
Figure 1. Effect of roasting temperature on leaching efficiency of Sn and Si (Time: 60 min).
Minerals 09 00605 g001
Figure 2. Effect of roasting time on leaching efficiency of Sn and Si (Temperature: 875 °C).
Figure 2. Effect of roasting time on leaching efficiency of Sn and Si (Temperature: 875 °C).
Minerals 09 00605 g002
Figure 3. Distribution of Sn and Si elements in the whole experimental flowsheet (under the most suitable conditions).
Figure 3. Distribution of Sn and Si elements in the whole experimental flowsheet (under the most suitable conditions).
Minerals 09 00605 g003
Figure 4. X-ray diffraction pattern of raw material.
Figure 4. X-ray diffraction pattern of raw material.
Minerals 09 00605 g004
Figure 5. XRD pattern of the roasted samples (Sample 1# in Figure 3).
Figure 5. XRD pattern of the roasted samples (Sample 1# in Figure 3).
Minerals 09 00605 g005
Figure 6. XRD pattern of the leach residues (Sample 2# in Figure 3).
Figure 6. XRD pattern of the leach residues (Sample 2# in Figure 3).
Minerals 09 00605 g006
Figure 7. Mole fractions of metasilicic and stannum species at different pH values.
Figure 7. Mole fractions of metasilicic and stannum species at different pH values.
Minerals 09 00605 g007
Figure 8. XRD patterns of the samples with different SiO2/SnO2 mole ratio (Temperature: 875 oC; Time: 60 min).
Figure 8. XRD patterns of the samples with different SiO2/SnO2 mole ratio (Temperature: 875 oC; Time: 60 min).
Minerals 09 00605 g008
Figure 9. XRD patterns of the samples roasted at different temperatures (Time: 60 min; SiO2/SnO2 molar ratio: 7:1).
Figure 9. XRD patterns of the samples roasted at different temperatures (Time: 60 min; SiO2/SnO2 molar ratio: 7:1).
Minerals 09 00605 g009
Figure 10. XRD pattern of the synthetic Na8SnSi6O18.
Figure 10. XRD pattern of the synthetic Na8SnSi6O18.
Minerals 09 00605 g010
Figure 11. The leaching efficiency Si and Sn of the synthetic Na8SnSi6O18. (a) leaching efficiency Si and Sn, (b) XRD pattern of leach residues.
Figure 11. The leaching efficiency Si and Sn of the synthetic Na8SnSi6O18. (a) leaching efficiency Si and Sn, (b) XRD pattern of leach residues.
Minerals 09 00605 g011
Table 1. Chemical compositions of the original and retreated cassiterite concentrates (wt. %).
Table 1. Chemical compositions of the original and retreated cassiterite concentrates (wt. %).
ElementSnSiFeCaOSAl2O3ZnAsPb
Raw material42.93.98.868.315.111.161.210.500.38
Pretreated62.93.70.110.170.040.280.020.030.03
Table 2. Reaction constants for equilibrium reactions of Si(IV) and Sn(IV) species.
Table 2. Reaction constants for equilibrium reactions of Si(IV) and Sn(IV) species.
EquationReaction EquationEquilibrium Constants (K)
(4)Sn4+ + H2O = Sn(OH) 3+ + H+Kα1 = 103.73
(5)Sn4+ + 2H2O = Sn(OH)22+ + 2H+Kα1 = 101.29
(6)Sn4+ + 3H2O = Sn(OH)3+ + 3H+Kα1 = 100.47
(7)Sn4+ + 4H2O = Sn(OH)4 + 4H+Kα1 = 100.4
(8)Sn4+ + 5H2O = Sn(OH)5 + 5H+Kα2 = 10−7.7
(9)Sn4+ + 6H2O = Sn(OH)62− + 6H+Kα3 = 10−18.1
(10)SiO32− + H+ = HSiO3Kβ1 = 10−11.82
(11)HSiO3 + H+ = H2SiO3Kβ2 = 10−9.69

Share and Cite

MDPI and ACS Style

Zhang, Y.; Han, B.; Su, Z.; Chen, X.; Lu, M.; Liu, S.; Liu, J.; Jiang, T. Effect of Quartz on the Preparation of Sodium Stannate from Cassiterite Concentrates by Soda Roasting Process. Minerals 2019, 9, 605. https://doi.org/10.3390/min9100605

AMA Style

Zhang Y, Han B, Su Z, Chen X, Lu M, Liu S, Liu J, Jiang T. Effect of Quartz on the Preparation of Sodium Stannate from Cassiterite Concentrates by Soda Roasting Process. Minerals. 2019; 9(10):605. https://doi.org/10.3390/min9100605

Chicago/Turabian Style

Zhang, Yuanbo, Benlai Han, Zijian Su, Xijun Chen, Manman Lu, Shuo Liu, Jicheng Liu, and Tao Jiang. 2019. "Effect of Quartz on the Preparation of Sodium Stannate from Cassiterite Concentrates by Soda Roasting Process" Minerals 9, no. 10: 605. https://doi.org/10.3390/min9100605

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