Next Article in Journal
The Suppression of Hump Instability inside a Pump Turbine in Pump Mode Using Water Injection Control
Next Article in Special Issue
Development of a Process to Recycle NdFeB Permanent Magnets Based on the CaO-Al2O3-Nd2O3 Slag System
Previous Article in Journal
Numerical and Experimental Simulation of Hydraulic Fracture Propagation Mechanism in Conglomerate Formation Based on Hybrid Finite-Discrete Element Method
Previous Article in Special Issue
Assessment of Mass and Energy Balance of Waste Printed Circuit Board Recycling through Hydrogen Reduction in Black Copper Smelting Process
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Two Potential Ways of Vanadium Extraction from Thin Film Steelmaking Slags

1
School for Engineering of Matter, Transport and Energy, Arizona State University, Tempe, AZ 85287, USA
2
Institute of Iron and Steel Technology, Technische Universität Bergakademie Freiberg, 09599 Freiberg, Germany
*
Author to whom correspondence should be addressed.
Processes 2023, 11(6), 1646; https://doi.org/10.3390/pr11061646
Submission received: 1 April 2023 / Revised: 19 May 2023 / Accepted: 25 May 2023 / Published: 28 May 2023

Abstract

:
During the steelmaking process, a great amount of slag is generated as a by-product. Vanadium-bearing steelmaking slags are classified as hazardous and require special handling and storage due to the toxicity of vanadium pentoxides. At the same time, such slags are valuable sources for the recovery of vanadium. The present work reviews the investigations on vanadium recovery from CaO-SiO2-FeO-V2O5 thin film slags under the neutral and oxidizing conditions in the temperature range 1653 K to 1693 K (1380 °C to 1420 °C) using Single Hot Thermocouple Technique (SHTT). The slag samples were analyzed by SEM/EDX. The results indicated that vanadium pentoxide evaporation can be up to 17.73% under an oxidizing atmosphere, while spinel formation under an argon atmosphere was detected in the conditions of thin film slags.

1. Introduction

The conversion of existing linear economics to more a circular one, where the value of resources and materials is maintained for as long as possible and the waste generation is minimized, is important for developing a resource-efficient, sustainable and competitive economy.
Significant amounts of by-products are generated every year all over the world during the steel production process. In general, the manufacture of one metric ton of steel yields between 200 kg (for electric arc furnace) and 400 kg (for blast furnace and basic oxygen furnace) of by-products, such as slag, dust, and sludge [1]. Slag is the main by-product of the steelmaking industry. Apart from the main slagging components (e.g., CaO, SiO2, Al2O3, MgO, etc.), a significant amount of alloying elements appears in the slag (Mn, V, Cr, etc.) in oxidized form causing problems in further utilization of such slags, e.g., in civil engineering. Vanadium-containing slag is among them. It is classified as toxic material [2] due to its potential environmental and health risks. As a by-product of steelmaking, it requires proper management since it requires special handling, storage, and partial landfill. The utilization of vanadium-containing slag can contribute to the transition towards a more circular economy by preventing landfill waste and preserving natural resources, hence also reducing greenhouse gas (GHG) emissions. Moreover, vanadium is a very valuable alloying component of steelmaking, which significantly improves the quality of the final steel products by reducing the steel grain size, thus improving the wear resistance of the final products. Vanadium is also a key component in non-ferrous alloys, e.g., in Ti-6%Al-4%V which is used for the aircraft industry where lightness and high strength are important, and in additive manufacturing powder and for the production of body implants and prosthetics.
Existing technologies to remove vanadium from the slags—direct acid leaching [3,4,5,6,7,8], sodium salt roasting [9,10,11], hot slag reduction [12], and bioleaching [13]—have their advantages and limitations, and the choice of technology depends on different factors including economic and environmental considerations. The sodium roasting-water leaching method is the most commonly used one to extract vanadium from V-bearing slags [14], but at the same time it presents several environmental challenges, including (1) decomposition of sodium salts creates harmful gases such as HCl, Cl2, SO2, SO3, and CO2 [15], which pollute the air and damage equipment used in the extraction process; (2) wastewater generated during extraction has high concentrations of sodium and ammonium sulfates, making it difficult to reuse [16]; (3) the resulting leaching residues have a high sodium content, making them unusable and unsuitable for recycling because sodium corrodes the lining of blast furnaces and converters; (4) water leaching residue contains 1–3% of V2O5 which is currently turning into waste [17]. This results in the accumulation of leaching residues, which consume valuable land resources [18]. Bioleaching is based on the capability of certain microorganisms to transform solid compounds into extractable, water-soluble forms. However, it also has drawbacks including (1) bacterial leaching occurs very slowly (often taking several months) and cannot be quickly halted once initiated, unlike other methods; (2) the heat produced during the dissolving process can be fatal to the microorganisms involved; (3) large open spaces for treatment procedures are required; (4) high risk of contamination.
One of the ways to extract the compound from the liquid mix could be vaporization. It is well known that some slag constituents have higher vapor pressure in the molten state than the rest of the slag [19]. For example, much attention was dedicated to CaF2 evaporation from the slags [20,21,22]. Some pioneering works on CrO3 [23,24] and V2O5 [25,26] volatilization from steelmaking slags are relatively well addressed in the literature. These oxides constitute a serious health hazard, and their evaporation should be controlled in the areas of molten steel tapping. Vanadium pentoxide evaporation leads to a vanadium loss during steel alloying with V2O5 [27]. Additionally, a vaporization study [28] for bulk synthetic slags with 6–8% V2O5 performed by two of the present authors should be mentioned. A combination of oxidation blowing and vacuum extirpation was applied to the 22 g of molten slag samples. After 60 min oxygen treatment time more than 77% of V2O5 was evaporated.
In this context, the present study aims to investigate the potential of extraction of vanadium from steelmaking slags through pyrometallurgical processes. The results of this study could provide valuable insights into the development of sustainable practices for the management of vanadium-containing slag and the recovery of critical metals from industrial by-products.

2. Thermodynamic Considerations

Figure 1 displays the stability of vanadium oxides under various oxygen partial pressures and temperatures. The results indicate that V2O5 remains stable across the entire temperature range under oxygen partial pressures equivalent to atmospheric conditions. Consequently, vanadium metal and lower vanadium oxides tend to oxidize to V2O5 upon exposure to air, even if the process is gradual. Moreover, the lower oxides of vanadium exhibit stability with increasing temperature and decreasing oxygen partial pressure.
The phase stability diagram of the VOx system was assessed utilizing FactSage [31], and the outcomes are illustrated in Figure 2. It can be seen from the diagram that V2O3 can remain solid at low oxygen pressures while VO2 is mainly in a liquid state. V2O5 is the stable phase under elevated oxygen pressures.
The vapor pressure of vanadium pentoxide as a function of temperature (Figure 3) was calculated using the numerical-value equation given by Kubaschewski and Alcock [32]:
log P = 7100 T + 5.05
where P is vapor pressure in mmHg and T is the temperature in Kelvin.
Considering the relatively low melting point of V2O5 equal to 690 °C, one can assume that its vapor pressure would be elevated at temperatures above 1100 °C. Consequently, some vanadium most probably is lost as pentoxide vapor even during the salt roasting process. A study conducted in 2012 [33] examined the evaporation of pure V2O5 under various atmospheres at temperatures ranging from 1450–1600 °C using thermogravimetry analysis (TGA). The results (Figure 4) showed that when exposed to a pure oxygen atmosphere (with a flow rate of 400 NmL/min), around 28% of a 100 mg sample could evaporate in 1 h 1600 °C, and in ca., 6 h nearly all of the sample was evaporated. In the air atmosphere vanadium pentoxide evaporation rate was considerably slower—it volatilized only on ca 80% after about 7 h. In the neutral (CO2) atmosphere ca 35% was evaporated in almost 7 h.
Wang et al. [27] studied the volatilization of pure V2O5 and its mixes in an argon atmosphere and reported a 4.34% mass loss for pure vanadium pentoxide after 10 min at 1600 °C.
In the liquid slag, vanadium can exist in two oxidation states: V3+ and V4+ [34]. At low oxygen levels, trivalent vanadium is the dominant form in the slag. To evaporate, the lower oxidation state vanadium oxides must first be oxidized to V2O5. This oxidation process can be represented by the following reaction:
V 2 O 3 + O 2 = V 2 O 5
This reaction is thermodynamically possible with ΔGo = −77.038 kJ/mol at a temperature of 1600 °C and O2 pressure of 101.3 kPa [35].
Authors [36,37,38] studied the effect of oxygen partial pressure and slag basicity on the vanadium oxidation state. The results suggest that the oxidation of V3+ in the solution state through reaction) is challenging. Similarly, V4+ can also be oxidized through this mechanism. To facilitate the evaporation of vanadium pentoxide from the liquid slag surface, it is necessary to oxidize the lower valent states of vanadium to V2O5 and transport it to the surface of the liquid slag. The main objective of the current study is to investigate the feasibility of removing vanadium in the form of oxide vapor from the liquid slag and subsequently condensing it for further use.

3. Experimental

3.1. Materials and Sample Preparation

For the experiments, the synthetic slag samples were prepared from pure oxides. CaCO3 powder was calcined at 1273 K (1000 °C) in a muffle furnace for 8 h to release CO2 gas from limestone decomposition. SiO2 and V2O5 powders were heat-treated at 383 K (110 °C) for 5 h to remove any moisture. The wustite (FeO) was synthesized from stoichiometric amounts of Fe and Fe2O3 powders. The mixture was mixed in an agate mortar, placed into an iron crucible (purity 99.9%) with a tight iron lid, and kept at 1373 K (1100 °C) in an argon atmosphere for 24 h to reach equilibrium. The crucible was then air-cooled. The FeO produced was examined by XRD to confirm the absence of metallic iron and magnetite.
After mixing the chemicals in appropriate proportions, the slag was pre-melted in an induction furnace in an inert atmosphere in an iron crucible at 1673 K (1400 °C), sufficient to get homogeneous liquid slag. Then the crucible with a sample was fast-cooled without contact with air. The slag composition by XRF was found to be close to the fractions from the initially weighted amounts. No vanadium loss was detected.
For each slag composition, a series of experimental samples was prepared. This was carried out to check the influence of different treatment time sequences under oxygen and argon atmosphere. The prepared series of slag mixtures were stored in a desiccator until used.
The following slag compositions were used for the experiments (Table 1). This slag composition was already tested by authors previously [25] but the method used for the evaluation of the results was different. Moreover. In the current work, authors focused attention on the influence of basicity as well as studying the effect of atmospheres used for heat treatment.

3.2. Experimental Set-Up

The experiments were performed using Single Hot Thermocouple Technique (SHTT) (Figure 5) custom built at TU Bergakademie Freiberg, Germany. SHTT is usually used for studying the crystallization and the development of time-temperature-transformation diagrams. As thin film experiments at SHTT have perfect conditions to study the evaporation process due to a very high surface-to-volume ratio, it was used for the current experiments to have both slag surfaces available for evaporation.
A thin film sample of slag weighing approximately 5 milligrams was held on a B-type loop thermocouple made of Pt-30 wt.% Rh/Pt-6 wt.% Rh wire with a diameter of 0.5 mm. The loop size was approximately 2.5 mm in outer diameter. The thermocouple was utilized both for heating the sample and measuring its temperature. The temperature was controlled by a temperature controller. The experiments were carried out within a temperature range of 1380 °C to 1420 °C.
The heat treatment procedures applied in this study consisted of two sequences. The first sequence involved heating the sample at a rate of 100 °C/min in an argon atmosphere to the target temperature, followed by a pre-treatment period of 35 min to establish thermal equilibrium. The sample was then air treated for varying durations (0, 1, 5, 20, 45, and 60 min) and then cooled down at a rate of 40 °C/min in argon. The second sequence was similar, with the exception that the sample was treated with argon instead of air for the same duration periods.

3.3. Samples Analysis

After the SHTT experiments, the samples were mounted and analyzed on a Scanning Electron Microscope (SEM) coupled with energy-dispersive X-ray spectroscopy (EDX). The used SEM was a Quanta 250 FEG manufactured by FEI, and the EDX analysis was performed using an Octane Plus silicon drift detector (SDD) from Ametek/EDAX. To determine the elemental composition of each slag sample, their polished and well-defined zone were analyzed using both EDX area analysis and elemental mapping. Elemental mapping was found to be more precise in terms of a relative measuring error and was therefore chosen for further evaluation. Unlike backscatter electron (BSE) images that have a limited gray-scale range, EDX mapping displays the true spatial distribution of each element based on the time the beam dwells on each point. The EDX mapping parameters were set to a mapping resolution of 512 × 400 points, 32 frames, and a dwell time of 1000 s. Bulk chemical analysis by XRF after the thin slag experiments was not possible due to the small size of the samples, so the average composition was determined with EDX. V2O5 content was calculated stoichiometrically by dividing vanadium content in the sample by a conversion factor of 0.56 (assuming that all vanadium is in the form of V2O5).

4. Results and Discussion

Figure 6 displays a cross-sectional view of the thin slag formed in the thermocouple loop. The convex shape of the slag film is due to the competing effect of wetting between the Pt-thermocouple and the slag’s surface tension. Such a convex shape has the advantage of increasing the evaporative surface area exposed to the surrounding atmosphere.
In the previous thin-film slag studies conducted by the authors [28,29], simple spot analysis and line SEM-EDX analyses to determine the vanadium content in the slag sample were performed. This method involved analyzing 10 points on each sample and calculating the average vanadium value. However, after the application of the mapping technique such analysis was found to underestimate the vanadium content in the slag. To address this issue, area mapping was employed, which confirmed the loss of vanadium from the sample, but to a much lesser extent. Figure 7 compares the average vanadium pentoxide content in the slag obtained through the previous method (a) with those obtained through area analysis (b) as a function of oxygen treatment time.
The investigation of four slag systems with varying binary basicities (CaO/SiO2 = 0.8, 1.0, 1.2, 1.4) revealed that vanadium pentoxide evaporation rate as a function of air treatment time is higher at lower basicities of the slag (Figure 8). The vanadium pentoxide extraction rate for the slag with binary basicity of 0.8 was found to be 17.73%. In comparison, the basicities of 1.2 and 1.4 showed a vanadium pentoxide extraction rate of 10.45%. It is hypothesized that the evaporation of V is primarily controlled by diffusion through the slag mass.
Calculation of viscosity (Figure 9a) for the initial samples was carried out with the use of the Iida model [39], given in tables by Mills [40]. Iida viscosity model considers the network structure of the slag using modified basicity index Bi*:
μ = A μ 0 exp E B i * ,   ( Pa * s )
where μ 0 is the viscosity of hypothetical network forming melt and B i * is a modified basicity index.
A = 1.745 1.962 × 10 3 T + 7 × 10 7 T 2
E = 11.11 3.65 × 10 3 T
B i * = α i % i B α i % i A
where A—acid oxides and B—basic oxides or fluorides.
μ 0 = i = 1 n μ 0 i X i
and μ 0 i values can be calculated using the expression:
μ 0 i = 1.8 × 10 7 M i T m i 1 / 2 exp H i / R T V m i 2 / 3 exp H i / R T m i ,   ( Pa · s )
H i = 5.1 T m i 1 / 2
where M is the formula weight, Vm is the molar volume at the melting point Tm, R is the gas constant, X is the mole fraction, and i refers to the component.
The surface tension was calculated using a partial molar approach (Figure 9b) [40]:
γ = ΣX1 γ1 + X2 γ2 + X3 γ3 + …
Despite the absence of V2O5 data in Mills’ table, it was presumed that the vanadium present in the liquid existed in the form of V3+ or V4+ ions. The vanadium content was then converted into V2O3. In the calculation of the optical basicity of the slag, it was postulated that V2O3 possessed similar characteristics to Cr2O3. This assumption was made based on the similar effect on viscosity [41,42] and surface tension of both chromium and vanadium oxides [43]. Analysis revealed a decrease in viscosity along with the basicity rise. At the same time, surface tension (Figure 9b) is slightly increasing with basicity. Therefore, the physical properties of liquid should generally favor the evaporation process. However, this contradicts our observations and indicates other phenomena affecting the evaporation of V2O5.
The most probable explanation of the inhibitory effect of basicity on the evaporation rate could be the binding of V2O5 with CaO. Wang et al. [27] studied the effect of CaO on V2O5 volatilization and reported that the addition of CaO reduced the activity of V2O5 and decreased its volatilization rate. The reason is that CaO and V2O5 form calcium vanadates—namely CaO.V2O5, 2CaO.V2O5, and 3CaO.V2O5. These are relatively high melting point compounds (Figure 10). On the other hand, according to the FactSage calculations, the amount of liquid phase in the studied slag systems is also affected by the increase of basicity. At 1400 °C, the slag system with 44.1 wt.% CaO, 31.5 wt.% SiO2, 20 wt.% FeO, 4.4 wt.% V2O5 and the binary basicity of 1.4 has 13.4% of solid fraction while other slags are completely liquid. As per the three-phase diagram for CaO-SiO2-FeO normalized without V2O5 at 1400 °C (Figure 11), all four slags should be in a liquid state.
Thus, it is expected that the evaporation rate of vanadium pentoxide would be significantly lower in samples with a higher basicity, as the fraction of the solid phase increases together with calcium vanadate formation.
Wang et al. [27] studied the effect of CaO on V2O5 volatilization and discovered that the addition of CaO reduced the activity of V2O5 and decreased its volatilization rate.
The slag sample with the binary basicity of 0.8 had homogeneous dendritic precipitations after 1 min air treatment. After 60 min only some dendritic precipitates were left on the sample’s borders. The rest of the sample was constituted of an amorphous matrix (Figure 12). Spot analysis of the sample after 1 min treatment at the air conditions revealed a high concentration of V in the matrix, while dendrites had an elevated share of Fe.
The samples with the binary basicities of 1.0, 1.2, and 1.4 had similar behavior during air heat treatment. After 1 min slag had an amorphous matrix surrounded by iron oxide dendrites and V-rich oxides bound into calcium silicate Ca2SiO4 matrix (Figure 13). As can be seen from the figure, the size of the crystals was changing while increasing the treatment time. The crystal formations, in particular vanadium ferrites FeV2O4 and Fe2VO4, for CaO-SiO2-FeO- V2O5 system were studied by Semykina et al. [44].
To assure that no vanadium from the slag was alloyed with the Pt-thermocouple during the SHTT experiments, the thermocouple cross-section was examined. The SEM/EDX analysis revealed no vanadium in the Pt phase (Figure 14).
During the experiments conducted under a pure argon atmosphere, a high amount of bubbles formation was detected (Figure 15). The bubbles are not colliding with other bubbles due to low interfacial tension and high viscosity. No loss of vanadium was detected by the SEM/EDX mapping.
The atomic mass of different elements on Spot 1 indicated that vanadium existed in the form of spinel (FeV2O4) (Figure 16 and Figure 17). SEM/EDX mapping of elements showed that the highest V content was in the inner surfaces of the bubbles. This indicates that in the absence of an oxygen atmosphere, vanadium pentoxide remains entrapped in the bubbles formed inside the slag.
Scientists have shown a keen interest in FeV2O4 due to the presence of various quantum mechanical phenomena, including successive structural phase transitions, orbital degree of freedom for both Fe2+ and V3+ ions, spin frustration while interacting between electromagnetism and lattice structure, and spin glass [45,46]. In the current study, FeV2O4 crystals were observed under an argon atmosphere during the formation of bubbles in the SHTT equipment. The other authors [44] have previously obtained similar results for V-bearing slags during experiments in a horizontal furnace under an argon atmosphere and oxygen partial pressures P O 2 = 10−1 − 10−2 Pa. The obtained spinels could potentially be separated from the slag using magnetic separation.

5. Conclusions

During the experiments with thin films of vanadium-containing slags at 1380–1420 °C in argon and air atmospheres different phenomena of vanadium oxides behavior were observed. Under the air atmosphere vanadium oxides tend to evaporate from the slag. The maximum achieved evaporation was 17.73% after 1 h of air treatment. An increase in basicity lowers the evaporation rate due to the formation of calcium vanadates. Under the argon atmosphere, slag had a high number of bubbles entrapped. SEM/EDX results revealed FeV2O4 spinel formation which is valuable material as such and can be magnetically separated.
The obtained results can serve as the background for two ways of V-extraction and as a background for designing a new vanadium extraction technique on the plant practice. The synthetic slags from the present study were exposed only to the natural convection of air and argon. It is expected that targeted air or oxygen treatment should improve the oxidation and evaporation of vanadium. Additionally, results raise a concern about the potential health hazard of working with the molten vanadium-bearing slags associated with the inhalation of toxic V2O5 vapors.

Author Contributions

Conceptualization, T.S., O.V. and S.S.; methodology, T.S.; software, T.S. and L.N.; validation, T.S., L.N., O.V. and S.S.; formal analysis, T.S.; investigation, T.S.; resources, T.S. and L.N.; data curation, O.V. and S.S.; writing—original draft preparation, T.S.; writing—review and editing, O.V., L.N. and S.S.; supervision, O.V. and S.S.; funding acquisition, T.S. All authors have read and agreed to the published version of the manuscript.

Funding

Financial support to Shyrokykh by European Social Fund (ESF) for this research is greatly acknowledged.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

The authors are thankful to the Institute for Iron and Steel Technology (IEST), TU Bergakademie Freiberg for making SHTT experimental facilities available and Institute of Energy Process Engineering and Chemical Engineering (IEC) for great help with SEM/EDX analysis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Worldsteel Association Fact Sheet. Steel Industry Co-Products. Available online: https://worldsteel.org/wp-content/uploads/Fact-sheet-Steel-industry-co-products.pdf (accessed on 1 March 2023).
  2. IARC Working Group on the Evaluation of Carcinogenic Risks to Humans. Cobalt in Hard Metals and Cobalt Sulfate, Gallium Arsenide, Indium Phosphide, and Vanadium Pentoxide; World Health Organization International Agency for Research on Cancer: Lyon, France; Geneva, Switzerland, 2006; ISBN 978-92-832-1286-7. [Google Scholar]
  3. Bin Li, X.; Wei, C.; Deng, Z.G.; Li, M.T.; Li, C.X.; Xu, H.S. Acid Leaching of Vanadium from a Vanadium Residue. Adv. Mater. Res. 2011, 402, 243–248. [Google Scholar] [CrossRef]
  4. Xiang, J.; Huang, Q.; Lv, X.; Bai, C. Extraction of vanadium from converter slag by two-step sulfuric acid leaching process. J. Clean. Prod. 2018, 170, 1089–1101. [Google Scholar] [CrossRef]
  5. Zhang, G.-Q.; Zhang, T.-A.; Lü, G.-Z.; Zhang, Y.; Liu, Y.; Liu, Z.-L. Extraction of vanadium from vanadium slag by high pressure oxidative acid leaching. Int. J. Miner. Met. Mater. 2015, 22, 21–26. [Google Scholar] [CrossRef]
  6. Aarabi-Karasgani, M.; Rashchi, F.; Mostoufi, N.; Vahidi, E. Leaching of vanadium from LD converter slag using sulfuric acid. Hydrometallurgy 2010, 102, 14–21. [Google Scholar] [CrossRef]
  7. Zhang, X.; Fang, D.; Song, S.; Cheng, G.; Xue, X. Selective leaching of vanadium over iron from vanadium slag. J. Hazard. Mater. 2019, 368, 300–307. [Google Scholar] [CrossRef] [PubMed]
  8. Li, H.-Y.; Wang, C.-J.; Yuan, Y.-H.; Guo, Y.; Diao, J.; Xie, B. Magnesiation roasting-acid leaching: A zero-discharge method for vanadium extraction from vanadium slag. J. Clean. Prod. 2020, 260, 121091. [Google Scholar] [CrossRef]
  9. Chen, D.; Zhao, L.; Liu, Y.; Qi, T.; Wang, J.; Wang, L. A novel process for recovery of iron, titanium, and vanadium from titanomagnetite concentrates: NaOH molten salt roasting and water leaching processes. J. Hazard. Mater. 2013, 244-245, 588–595. [Google Scholar] [CrossRef]
  10. Liu, B.; Meng, L.; Zheng, S.; Li, M.; Wang, S. A novel method to extract vanadium from high-grade vanadium slag: Non-salt roasting and alkaline leaching. Physicochem. Probl. Miner. Process. 2018, 54, 657–667. [Google Scholar]
  11. Li, H.-Y.; Fang, H.-X.; Wang, K.; Zhou, W.; Yang, Z.; Yan, X.-M.; Ge, W.-S.; Li, Q.-W.; Xie, B. Asynchronous extraction of vanadium and chromium from vanadium slag by stepwise sodium roasting–water leaching. Hydrometallurgy 2015, 156, 124–135. [Google Scholar] [CrossRef]
  12. Ye, G.; Burstrom, E.; Kuhn, M.; Piret, J. Reduction of steel-making slags for recovery of valuable metals and oxide materials. Scand. J. Met. 2003, 32, 7–14. [Google Scholar] [CrossRef]
  13. Mirazimi, S.M.J.; Rashchi, F. Optimization of Bioleaching of a Vanadium Containing Slag Using RSM. In Proceedings of the 7th International Chemical Engineering Congress & Exihibition, Kish, Iran, 21–24 November 2011; p. 11. [Google Scholar]
  14. Mahdavian, A.; Shafyei, A.; Alamdari, E.K.; Haghshenas, D.F. Recovery of vanadium from esfahan steel company steel slag; Optimizing of roasting and leaching parameters. Int. J. Iron Steel Soc. Iran 2006, 3, 17–21. [Google Scholar]
  15. Jia, L.; Zhang, Y.; Tao, L.; Jing, H.; Bao, S. A methodology for assessing cleaner production in the vanadium extraction industry. J. Clean. Prod. 2014, 84, 598–605. [Google Scholar] [CrossRef]
  16. Yuan, J.; Cao, Y.; Fan, G.; Du, H.; Dreisinger, D.; Han, G.; Li, M. Study on the Mechanisms for Vanadium Phases’ Transformation of Vanadium Slag Non-Salt Roasting Process. In Proceedings of the Rare Metal Technology 2020; Azimi, G., Forsberg, K., Ouchi, T., Kim, H., Alam, S., Baba, A.A., Eds.; Springer International Publishing: Cham, Switzerland, 2020; pp. 235–242. [Google Scholar]
  17. Hu, P.; Zhang, Y.; Liu, H.; Liu, T.; Li, S.; Zhang, R.; Guo, Z. High efficient vanadium extraction from vanadium slag using an enhanced acid leaching-coprecipitation process. Sep. Purif. Technol. 2023, 304, 122319. [Google Scholar] [CrossRef]
  18. Ji, Y.; Shen, S.; Liu, J.; Yan, S.; Zhang, Z. Green and Efficient Process for Extracting Chromium from Vanadium Slag by an Innovative Three-Phase Roasting Reaction. ACS Sustain. Chem. Eng. 2017, 5, 6008–6015. [Google Scholar] [CrossRef]
  19. Ling, H.; Malfliet, A.; Blanpain, B.; Guo, M. Evaporation of Antimony Trioxide from Antimony Slag by Nitrogen Injection in a Top-Submerged Lance Smelting Set-Up. In Proceedings of the 12th International Symposium on High-Temperature Metallurgical Processing; Peng, Z., Hwang, J.-Y., White, J.F., Downey, J.P., Gregurek, D., Zhao, B., Yücel, O., Keskinkilic, E., Jiang, T., Mahmoud, M.M., Eds.; Springer International Publishing: Cham, Switzerland, 2022; pp. 133–142. [Google Scholar]
  20. Persson, M.; Seetharaman, S.; Seetharaman, S. Kinetic Studies of Fluoride Evaporation from Slags. ISIJ Int. 2007, 47, 1711–1717. [Google Scholar] [CrossRef]
  21. Ju, J.-T.; Ji, G.-H.; Tang, C.-M.; An, J.-L. Fluoride Evaporation and Melting Characteristics of CaF2 -CaO-Al2 O3-MgO-Li2 O- (TiO2) Slag for Electroslag Remelting. Steel Res. Int. 2020, 91, 2000111. [Google Scholar] [CrossRef]
  22. Shimizu, K.; Cramb, A. Fluoride Evaporation from CaF2-SiO2-CaO Slags and Mold Fluxes in Dry and Humid Atmospheres. High Temp. Mater. Process. 2003, 22, 237–246. [Google Scholar] [CrossRef]
  23. Seetharaman, S.; Albertsson, G.J.; Scheller, P. Studies of Vaporization of Chromium from Thin Slag Films at Steelmaking Temperatures in Oxidizing Atmosphere. Met. Mater. Trans. B 2013, 44, 1280–1286. [Google Scholar] [CrossRef]
  24. Cheremisina, E.; Schenk, J. Chromium Stability in Steel Slags. Steel Res. Int. 2017, 88, 1700206. [Google Scholar] [CrossRef]
  25. Seetharaman, S.; Shyrokykh, T.; Schröder, C.; Scheller, P.R. Vaporization Studies from Slag Surfaces Using a Thin Film Technique. Met. Mater. Trans. B 2013, 44, 783–788. [Google Scholar] [CrossRef]
  26. Shyrokykh, T.; Schröder, C.; Scheller, P.R.; Shatokha, V.; Seetharaman, S. Studies of High Temperature Surface Oxidation of FeO-CaO-SiO2-V2O5 Slags with the Use of Single Hot Thermocouple Technique. Metall. Mater. Trans. B 2012, 44, 4. [Google Scholar]
  27. Wang, W.X.; Xue, Z.L.; Song, S.Q.; Li, P.; Chen, Z.C.; Liu, R.N.; Wang, G.L. Research on High-Temperature Volatilization Characteristics of V2O5 during Direct Alloying of Smelting Vanadium Steel. Adv. Mater. Res. 2012, 557–559, 182–186. [Google Scholar] [CrossRef]
  28. Shyrokykh, T.; Wei, X.; Seetharaman, S.; Volkova, O. Vaporization of Vanadium Pentoxide from CaO-SiO2-VOx Slags During Alumina Dissolution. Met. Mater. Trans. A 2021, 52, 1472–1483. [Google Scholar] [CrossRef]
  29. Bergerud, A.J. Phase Stability and Transformations in Vanadium Oxide Nanocrystals. Ph.D. Thesis, University of California, Berkeley, CA, USA, 2016. [Google Scholar]
  30. Yamaguchi, I.; Manabe, T.; Tsuchiya, T.; Nakajima, T.; Sohma, M.; Kumagai, T. Preparation and Characterization of Epitaxial VO2Films on Sapphire Using Postepitaxial Topotaxy Route via Epitaxial V2O3Films. Jpn. J. Appl. Phys. 2008, 47, 1022–1027. [Google Scholar] [CrossRef]
  31. FactSage, Version 7.2; Centre for Research in Computational Thermochemistry: Quebec, QC, Canada; GTT-Technologies: Herzogenrath, Germany.
  32. Kubaschewski, O.; Alcock, C.B. Metallurgical Thermochemistry, 5th ed.; Metallurgy: Moscow, Russia, 1982. [Google Scholar]
  33. Yang, Y.; Teng, L.; Seetharaman, S. Kinetic Studies on Evaporation of Liquid Vanadium Oxide, VOx (Where x = 4 or 5). Met. Mater. Trans. B 2012, 43, 1684–1691. [Google Scholar] [CrossRef]
  34. Wang, L.; Teng, L.; Chou, K.-C.; Seetharaman, S. Determination of Vanadium Valence State in CaO-MgO-Al2O3-SiO2 System By High-Temperature Mass Spectrometry. Met. Mater. Trans. B 2013, 44, 948–953. [Google Scholar] [CrossRef]
  35. HSC Chemistry, Version 9.0; Chemistry Software: Houston, TX, USA: Houston, TX, USA.
  36. Wang, H.; Wang, L.; Seetharaman, S. Determination of Vanadium Oxidation States in CaO-MgO-Al2O3-SiO2-VOxSystem by K Edge XANES Method. Steel Res. Int. 2015, 87, 199–209. [Google Scholar] [CrossRef]
  37. Mittelstadt, R.; Schwerdtfeger, K. The dependence of the oxidation state of vanadium on the oxygen pressure in melts of VOx, Na2O-VOx, and CaO-SiO2-VOx. Met. Trans. B 1990, 21, 111–120. [Google Scholar] [CrossRef]
  38. Farah, H.; Brungs, M. Oxidation-reduction equilibria of vanadium in CaO-SiO2, CaO-Al2O3-SiO2 and CaO-MgO-SiO2 melts. J. Mater. Sci. 2003, 38, 1885–1894. [Google Scholar] [CrossRef]
  39. Iida, T.; Sakai, H.; Kita, Y.; Shigeno, K. An Equation for Accurate Prediction of the Viscosities of Blast Furnace Type Slags from Chemical Composition. ISIJ Int. 2000, 40, S110–S114. [Google Scholar] [CrossRef]
  40. Mills, K. Ken Mills–Slag Modelling. Available online: https://www.pyrometallurgy.co.za/KenMills/ (accessed on 26 July 2020).
  41. Wang, Z.; Bai, J.; Kong, L.; Wen, X.; Li, X.; Bai, Z.; Li, W.; Shi, Y. Viscosity of coal ash slag containing vanadium and nickel. Fuel Process. Technol. 2014, 136, 25–33. [Google Scholar] [CrossRef]
  42. Wu, T.; Zhang, Y.; Yuan, F.; An, Z. Effects of the Cr2O3 Content on the Viscosity of CaO-SiO2-10 Pct Al2O3-Cr2O3 Quaternary Slag. Met. Mater. Trans. B 2018, 49, 1719–1731. [Google Scholar] [CrossRef]
  43. Mills, K. Slag Atlas, 2nd ed.; VDEh: Düsseldorf, Germany, 1995. [Google Scholar]
  44. Semykina, A.; Dzhebian, I.; Shatokha, V. On the Formation of Vanadium Ferrites in CaO-SiO2-FeO-V2O5 Slags. Steel Res. Int. 2012, 83, 1129–1134. [Google Scholar] [CrossRef]
  45. Katsufuji, T.; Suzuki, T.; Takei, H.; Shingu, M.; Kato, K.; Osaka, K.; Takata, M.; Sagayama, H.; Arima, T.-H. Structural and Magnetic Properties of Spinel FeV2O4with Two Ions Having Orbital Degrees of Freedom. J. Phys. Soc. Jpn. 2008, 77, 053708. [Google Scholar] [CrossRef]
  46. Myoung, B.R.; Kim, S.J.; Lim, J.T.; Kouh, T.; Kim, C.S. Microscopic evidence of magnetic and structure phase transition in multiferroic spinel FeV2O4. AIP Adv. 2017, 7, 055828. [Google Scholar] [CrossRef]
Figure 1. Ellingham diagram for vanadium oxides. Dashed lines show the oxides stability at P O 2 = 0.21 (air), 10−5, 10−10 atm [29,30].
Figure 1. Ellingham diagram for vanadium oxides. Dashed lines show the oxides stability at P O 2 = 0.21 (air), 10−5, 10−10 atm [29,30].
Processes 11 01646 g001
Figure 2. Phase stability diagram of VOx.
Figure 2. Phase stability diagram of VOx.
Processes 11 01646 g002
Figure 3. Vapor pressure of V2O5 in equilibrium with liquid V2O5 as a function of temperature.
Figure 3. Vapor pressure of V2O5 in equilibrium with liquid V2O5 as a function of temperature.
Processes 11 01646 g003
Figure 4. Dimensionless mass loss of V2O5 in oxygen and air atmosphere as a function of time and temperature [33].
Figure 4. Dimensionless mass loss of V2O5 in oxygen and air atmosphere as a function of time and temperature [33].
Processes 11 01646 g004
Figure 5. The experimental apparatus—Single Hot Thermocouple Technique (SHTT) [25] 1—vacuum chamber, 2—insert, 3—thermocouple, 4—sealed window, 5—electronic circuit, 6—CCD camera, 7—computer control, 8—lamp, 9—additional heating, 10—gas exit to pump, 11—gas inlet for air or Ar.
Figure 5. The experimental apparatus—Single Hot Thermocouple Technique (SHTT) [25] 1—vacuum chamber, 2—insert, 3—thermocouple, 4—sealed window, 5—electronic circuit, 6—CCD camera, 7—computer control, 8—lamp, 9—additional heating, 10—gas exit to pump, 11—gas inlet for air or Ar.
Processes 11 01646 g005
Figure 6. The cross-section of the slag film in the thermocouple loop.
Figure 6. The cross-section of the slag film in the thermocouple loop.
Processes 11 01646 g006
Figure 7. The average content of vanadium pentoxide as a function of oxygen treatment time was measured with spot analysis (a) [28] and with area mapping (b) in SEM EDX.
Figure 7. The average content of vanadium pentoxide as a function of oxygen treatment time was measured with spot analysis (a) [28] and with area mapping (b) in SEM EDX.
Processes 11 01646 g007
Figure 8. Effect of basicity on vanadium pentoxide evaporation rate as a function of air treatment time.
Figure 8. Effect of basicity on vanadium pentoxide evaporation rate as a function of air treatment time.
Processes 11 01646 g008
Figure 9. Viscosity (a) and surface tension (b) were calculated using the Iida model and method 2a in Mills tables [40].
Figure 9. Viscosity (a) and surface tension (b) were calculated using the Iida model and method 2a in Mills tables [40].
Processes 11 01646 g009
Figure 10. Phase diagram of CaO-V2O5 system [43].
Figure 10. Phase diagram of CaO-V2O5 system [43].
Processes 11 01646 g010
Figure 11. Phase diagram for CaO-SiO2-FeO system and compositions of the samples with CaO/SiO2 = 0.8 (point 1), 1.0 (point 2), 1.2 (point 3), and 1.4 (point 4) normalized without V2O5 at 1400 °C.
Figure 11. Phase diagram for CaO-SiO2-FeO system and compositions of the samples with CaO/SiO2 = 0.8 (point 1), 1.0 (point 2), 1.2 (point 3), and 1.4 (point 4) normalized without V2O5 at 1400 °C.
Processes 11 01646 g011
Figure 12. SEM micrographs of the slag sample 33.6% CaO—42.0% SiO2—20% FeO—4.4% V2O5 (CaO/ SiO2 = 0.8) at different air treatment time: (a) 1 min, (b) 60 min (T = 1400 °C).
Figure 12. SEM micrographs of the slag sample 33.6% CaO—42.0% SiO2—20% FeO—4.4% V2O5 (CaO/ SiO2 = 0.8) at different air treatment time: (a) 1 min, (b) 60 min (T = 1400 °C).
Processes 11 01646 g012
Figure 13. SEM micrographs of the slag sample 37.8% CaO—37.8% SiO2—20% FeO—4.4% V2O5 at different air treatment times: (a) 1 min, (b) 60 min (T = 1400 °C); Spot 1 (dark grey color)—wollastonite CaSiO3, Spot 2 (white color)—iron oxides, Free Draw (light grey color)—V-rich oxides (possibly FeV2O4) and calcium silicate Ca2SiO4.
Figure 13. SEM micrographs of the slag sample 37.8% CaO—37.8% SiO2—20% FeO—4.4% V2O5 at different air treatment times: (a) 1 min, (b) 60 min (T = 1400 °C); Spot 1 (dark grey color)—wollastonite CaSiO3, Spot 2 (white color)—iron oxides, Free Draw (light grey color)—V-rich oxides (possibly FeV2O4) and calcium silicate Ca2SiO4.
Processes 11 01646 g013
Figure 14. Spectral analysis of Pt-thermocouple by SEM/EDX and its spectrum.
Figure 14. Spectral analysis of Pt-thermocouple by SEM/EDX and its spectrum.
Processes 11 01646 g014
Figure 15. Molten slag inside the thermocouple during the SHTT experiment.
Figure 15. Molten slag inside the thermocouple during the SHTT experiment.
Processes 11 01646 g015
Figure 16. SEM micrograph of argon-treated slag sample.
Figure 16. SEM micrograph of argon-treated slag sample.
Processes 11 01646 g016
Figure 17. SEM/EDX results of argon-treated slag sample.
Figure 17. SEM/EDX results of argon-treated slag sample.
Processes 11 01646 g017
Table 1. Synthetic slag compositions.
Table 1. Synthetic slag compositions.
Sample NameCaO/SiO2
Ratio
CaO, wt.%SiO2, wt.%FeO, wt.%V2O5, wt.%
CSFV10.833.642.020.04.4
CSFV21.037.837.820.04.4
CSFV31.241.234.420.04.4
CSFV41.444.131.520.04.4
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Shyrokykh, T.; Neubert, L.; Volkova, O.; Sridhar, S. Two Potential Ways of Vanadium Extraction from Thin Film Steelmaking Slags. Processes 2023, 11, 1646. https://doi.org/10.3390/pr11061646

AMA Style

Shyrokykh T, Neubert L, Volkova O, Sridhar S. Two Potential Ways of Vanadium Extraction from Thin Film Steelmaking Slags. Processes. 2023; 11(6):1646. https://doi.org/10.3390/pr11061646

Chicago/Turabian Style

Shyrokykh, Tetiana, Lukas Neubert, Olena Volkova, and Seetharaman Sridhar. 2023. "Two Potential Ways of Vanadium Extraction from Thin Film Steelmaking Slags" Processes 11, no. 6: 1646. https://doi.org/10.3390/pr11061646

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