Next Article in Journal
Heap Leaching of Greek Low-Grade Nickel Oxide Ores by Dilute Sulphuric Acid at a Pilot-Plant Scale
Previous Article in Journal
ΣIDERWIN—A New Route for Iron Production
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Proceeding Paper

Production of Metallic Titanium by Electrowinning in Molten Salts of Titanium Oxycarbide Anode †

1
CEA, DES, ISEC, DMRC, University of Montpellier, Marcoule, 30207 Bagnols sur Ceze, France
2
IRCER, Institute of Research for Ceramics, University of Limoges, 87068 Limoges, France
3
LCTS, Laboratory of Thermo Structural Composites, University of Bordeaux, 33000 Bordeaux, France
*
Author to whom correspondence should be addressed.
Presented at International Conference on Raw Materials and Circular Economy, Athens, Greece, 5–9 September 2021.
Mater. Proc. 2021, 5(1), 63; https://doi.org/10.3390/materproc2021005063
Published: 7 December 2021
(This article belongs to the Proceedings of International Conference on Raw Materials and Circular Economy)

Abstract

:
The electrochemical behavior of Ti3+ in LiCl-LiF-TiF3 salt was investigated by cyclic and square wave voltammetries at 853 K. Both methods confirm the presence of a single reduction wave of Ti3+ ions to metal, at a potential of −2.3 V vs. Cl2/Cl. The closeness of the potentials of TiCxOy dissolution and Ti3+/Ti4+ wave is an issue during the electrorefining of the anode. A low current density has to be applied to stay within the titanium oxycarbide dissolution and avoid the formation of Ti4+. The titanium deposition was studied by electrorefining of a titanium metal plate in LiCl-LiF-TiF3 (0.62 mol/kg). The cathodic deposit analysis by XRD and SEM confirms the formation of titanium metal with an average grain size of 150 µm. The faradic deposition yields are above 85% and constant between 60 and 160 mA/cm2.

1. Introduction

Titanium, the ninth most abundant element in the earth’s crust, has many attractive properties such as lightweight, high specific strength, corrosion and oxidation resistance at elevated temperatures. It can be used as high-temperature structural material widely used in aerospace, petrochemical, automobile, and metallurgical industries [1].
This material is produced by the Kroll process since 1940s. The process allows the conversion of TiO2 into Ti metal. TiO2 is transformed into TiCl4 by the action of chlorine gas at high temperature (carbochlorination). TiCl4 is then converted to metal by reduction with metallic magnesium. This highly exothermic reaction leads to the formation of liquid MgCl2 and a metallic titanium sponge [2], according to the following reactions:
TiO 2 + Cl 2 ( g ) + C TiCl 4 ( g ) + CO 2 ( g )
TiCl 4 ( g ) + 2 Mg ( liq ) 2 MgCl 2 ( liq ) + Ti ( s )
Titanium sponge is finally obtained after a MgCl2 distillation step.
This high-cost batch [3] process requires optimal operation and wastes minimization, a chlorine gas production unit, and a magnesium metal production workshop by electrolysis of MgCl2 (resulting from the reduction reaction of TiCl4).
As an alternative method of the Kroll process, molten salt electrolysis has been frequently proposed.
The first difficulty in the electrochemistry of titanium lies in controlling its oxidation state in the molten salt solution [4].
This problem is especially true in molten chloride, for which Ti0, Ti2+, Ti3+, Ti4+ can exist in the salt. This particularity is explained by the existence of several equilibria that lead to losses in faradic efficiency [4]:
3 TiCl 4 + Ti 0 4 TiCl 3
2 TiCl 3 + Ti 0 3 TiCl 2
Since the 2000s, the FFC (Fray, Farthing, and Chen) [5] and OS (Ono and Suzuki) [6] processes have opened a new route through the electrolytic reduction of TiO2 into Ti in molten chlorides. The metal production at the cathode involves a reducing agent formed in situ by reduction of the molten salt solvent (e.g., calcium). These processes overcome the difficulty of multiple oxidation states of titanium in solution since titanium is present in the processes as porous TiO2 pellets reduced at the cathode. The TiO2 electroreduction leads to the formation of oxide ions that can be converted at a graphite anode into CO/CO2 or O2 with a non-consumable anode. However, the anodic reaction between graphite and oxides ions forms electro-reducible carbonates, which cause faradic yield losses and can contaminate the cathode deposit. Platinum, gold, SnO2 anodes cannot be considered as non-consumable anodes as degradation products have been frequently reported [7]. These processes therefore still require further developments to find a suitable anode, and the titanium product is often contaminated with carbon and residual oxygen [8].
The process studied in this work is the Chinuka process (also studied in the literature under the names MER and USTB). In this route, titanium oxycarbide (TiCxOy) is used as the anode in molten chlorides. When the right potential is applied, CO is released at the anode together with Ti3+ and/or Ti2+ ions that are dissolved into the molten salt and transferred to the cathode for electrodeposition [9,10].
In this work, the melt composition has been optimized to minimize the Ti2+ concentration in the salt phase. The solution adopted to solve the problem of multiple oxidation states consists in adjusting the degree of complexation of titanium ions in solution, either by modifying the so-called chloroacidity of the medium or by adding fluoride ions.
Numerous studies dealt with the electrochemical behavior of titanium in chloride medium. They allowed highlighting the presence of Ti3+ and Ti2+ and identifying a multistep reduction mechanism [11] according to the reactions:
Ti 4 + + e Ti 3 + ,   Ti 3 + + e Ti 2 +   and   Ti 2 + + 2 e Ti 0
The existence of these multiple oxidation states in solution involves disproportionation reactions (3) and (4), which can decrease the faradic efficiencies [12,13] and generate sludge formation [14].
In chloro-fluoride salts, studies [15] have shown that increasing the concentration of fluoride ions in the molten salt leads removing the Ti2+/Ti3+ electronic transition [16].
The overall idea of this work is to use a high concentration of fluoride ions (hence the use of a chloro-fluoride salt) in the Chinuka process.
To study the reduction mechanism of titanium ions, the choice was made on a salt rich in fluoride: LiCl-LiF (70–30 mol%).

2. Experimental

2.1. Preparation of Titanium Oxycarbide Anode

The carboreduction route has been used to synthesize titanium oxycarbide anode from a mixture of titanium dioxide (Alfa Aesar, purity 99.6%) and black carbon (Alfa Aesar, purity 99.9%). The raw powders were mixed in a mortar according to the stoichiometric proportion given by the following equation:
TiO 2 + 2 C   TiC 0.5 O 0.5 + 3 2 CO
The mixtures obtained were then heat treated under flowing argon (16 L·h−1) at 1773 K (1500 °C). This treatment lasted four hours followed by annealing at the same conditions as the first carboreduction in order to reach the desired stoichiometry of the titanium oxycarbide phase: Ti2CO [17]. Finally, the titanium oxycarbide powder was sintered by SPS (Spark Plasma Sintering) [18] at 100 MPa for 5 min at 1923 K (1650 °C) (Figure 1).

2.2. Electrolysis Process

The electrolyte consisted of a eutectic LiCl-LiF melt (70–30 mol%, 99–99.98%). It was dried at 423 K for a minimum of 24 h. Experiments were carried out in an argon glovebox. The salt was introduced in a glassy carbon crucible. Temperature was progressively risen to the working temperature (853–953 K) under argon atmosphere.
A concentration of 0.62 mol/kg of potassium hexafluorotitanate solute ( K 2 + , Ti 4 + , F 6 ) was added to the electrolyte and reduced to Ti3+ with a pure titanium plate. The working temperature was measured by a thermocouple protected by a glassy carbon tube and inserted into the melt.

2.3. Electrodes

To investigate the electrochemical behavior of titanium ions, a three-electrode setup was used. A tungsten wire with a diameter of 0.5 mm served as the working electrode, a molybdenum spiral was used as the counter electrode, and a molybdenum wire with a diameter of 1 mm was utilized as the quasi-reference electrode. All RE potentials were referred to Cl2/Cl redox couple. The conversion was obtained by graphical estimation of the potential at zero current in the linear variation region during the positive scan of voltammograms, as shown in reference [19].
For the electrorefining experiments, a titanium plate is used as an anode to study titanium electrodeposition. The cathode is a 5 mm diameter molybdenum rod.

2.4. Electrochemical Techniques Used

Electrochemical techniques, including cyclic voltammetry and square wave voltammetry, were used to study the mechanism of titanium ions reduction.

2.5. Materials Characterization

After electrolysis, the cathode was removed from the bath and washed in an ultrasonic tank containing a water solution to remove all residual salt. Titanium deposits were characterized using X-ray diffraction (XRD). The scanning electron microscopy (SEM) was carried out to observe the microstructure of titanium deposits.

3. Results and Discussion

3.1. Electrochemical Behavior of Titanium Ions

In order to be representative of an electrolyzer (reducing conditions), a titanium plate was dipped permanently into the molten salt.

3.1.1. Cyclic Voltammetry (CV)

The chloro-fluoride salt LiCl-LiF (70–30 mol%) has an F/Cl ratio of 0.43. After adding K2TiF6, the F/Cl ratio becomes 0.58 and the F/Ti ratio is 23.
The cyclic voltammograms obtained in LiCl-LiF containing dissolved titanium as well as the anodic polarization curve of titanium oxycarbide in LiCl-LiF at 853 K (580 °C) are shown in Figure 2. The scan rate was 100 mV/s.
After the addition of K2TiF6 and titanium plate in pure LiCl-LiF molten salt, two oxidation peaks associated with two reduction peaks appeared. The first one at around −0.5 V vs. Cl2/Cl corresponds to the Ti3+/Ti4+ soluble-soluble redox couple. The second one at about −2.3 V vs. Cl2/Cl is attributed to the reduction of Ti3+ in Ti metal in a single step associated to its sharp stripping peak:
Ti 3 + + 3   e Ti 0

3.1.2. Square Wave Voltammetry (SWV)

SWV was used to estimate the number of electrons exchanged on the −2.3 V vs. Cl2/Cl system. Figure 3 shows that the peak current density is increasing with the increase of frequency.
As shown in the inserted figure (in Figure 3), the differential current density (δJ (A·cm−2)) is proportional to the square root of the frequency (f (Hz)). Square wave voltammetry allows thus, by measuring the width at half-height of the peak (W1/2), to calculate the number of electrons exchanged, using the following equation [20]:
W 1 2 = 3.52 RT nF
where, R is the ideal gas constant (8.314 JK−1·mol−1), T is the temperature (K), n is the exchanging electron number, and F is the Faraday constant (96,500 C·mol−1).
To overcome the disturbance caused by nucleation phenomenon, the measurement of the half-width at half-height 0.5 W1/2 was performed on the second half of the metal reduction signal, as shown in Figure 3.
Using Equation (8), the average number of electrons in the frequency range of 9 to 49 Hz varies between 3.07 and 3.17. This confirms that the reduction peak at −2.3 V vs. Cl2/Cl can be attributed to the reduction of Ti3+ into metallic titanium.

3.1.3. Titanium Oxycarbide Anodic Dissolution

The red curve in Figure 2 shows the anodic polarization curve of TiCxOy. As the potential was raised beyond −1 V vs. Cl2/Cl, the current density increased rapidly, attributed to the occurrence of electrochemical dissolution of TiCxOy solid solution.
The dissolution potential of titanium oxycarbide is close to the potential of Ti3+/Ti4+ couple, i.e., −0.5 V. During the electrorefining of TiC0.5O0.5 low current densities will have to be applied in order to avoid the formation of Ti4+ and the problem of disproportionation reaction between Ti0 and Ti4+ that would decrease the faradic yield.

3.2. Electrodeposition of Titanium

Electrorefining of titanium was performed at high Ti3+ (0.62 mol/kg) concentration in LiCl-LiF at different temperatures. The elaboration of metallic titanium deposits by electrorefining in a molten LiCl-LiF salt is realized in the first step by anodic dissolution of a titanium plate in order to optimize the electrochemical parameters.
Electrolysis was performed at a current density of 60 mA·cm−2 for 2 h and 30 min. After electorefining tests, the cathode was coated with a mixture of titanium deposit and salt (see picture in Figure 4). The deposit was washed, dried, and weighed. The metal after cleaning is easier to obtain in a powder form than a dendritic one. Figure 4 shows XRD pattern of cathode deposit. It confirms that the obtained cathodic deposit is metallic titanium.
Figure 5 shows an SEM micrograph of a titanium deposit obtained through electrolysis, which exhibits a crystalline microstructure. The crystallites are approximately 150 µm, well above the limit where titanium is considered as pyrophoric (less than 5 µm [21]).
The faradic yields obtained from the calculated theoretical mass and the experimental mass obtained by weighing the deposits are about 85%.
The experiment was repeated at different current density over a range from 60 to 160 mA·cm−2, and temperature from 853 K (580 °C) to 953 K (700 °C). Faradic efficiencies were stable within this current range and slightly higher at 953 K.

4. Conclusions

The study of the reduction mechanism of titanium ions in LiCl-LiF chloro-fluoride salt showed that Ti3+ is reduced to Ti metal in a single step involving three electrons.
The dissolution potential of TiCxOy is close to the Ti3+/Ti4+ wave. In electrorefining of titanium oxycarbide, it is important to apply low current density to avoid the formation of Ti4+.
The faradic current efficiencies obtained after electrorefining a titanium plate between 60 and 160 mA·cm−2 are high, around 85% at 853 K and 90% at 953 K.
XRD and SEM characterization of the metal deposit shows the formation of metallic titanium with a grain size of around 150 µm.
Future work will be dedicated to the electrorefining of titanium oxycarbide in the experimental conditions optimized in this work.

References

  1. Yves COMBRES. Propriétés du Titane et de ses Alliages. Techniques de l’ingénieur, Réf.: M4780 V1. 10 March 2010. Available online: https://www.techniques-ingenieur.fr/res/pdf/encyclopedia/42357210-m4780.pdf (accessed on 4 December 2021).
  2. Kroll, W.; Fink, C.G.; Summers, D.B. The Production of Ductile Titanium. Trans. Electrochem. Soc. 1940, 78, 35–649. [Google Scholar] [CrossRef]
  3. van Vuuren, D.S. A critical evaluation of processes to produce primary titanium. J. S. Afr. Inst. Min. Metal. 2009, 109, 455–461. [Google Scholar]
  4. Marco, V.; Ginatta, G.T.T. Titanium Electrowinning. In Proceedings of the International Symposium on Ionic Liquids in Honour of Marcelle Gaune-Escard. Carry le Rouet, France, 27–28 June 2003; pp. 3–4. [Google Scholar]
  5. Chen, G.; Fray, D.J.; Farthing, T.W. Direct electrochemical reduction of titanium dioxide to titanium in molten calcium chloride. Nat. Cell Biol. 2000, 407, 361–364. [Google Scholar] [CrossRef] [PubMed]
  6. Ono, K.; Suzuki, R.O. A new concept for producing Ti sponge: Calciothermic reduction. J. Miner. Metals Mater. Soc. 2002, 54, 59–61. [Google Scholar] [CrossRef]
  7. Mohandas, K.S. Direct electrochemical conversion of metal oxides to metal by molten salt electrolysis: A review. Miner. Process. Extr. Met. 2013, 122, 195–212. [Google Scholar] [CrossRef]
  8. Bertolini, M.; Shaw, L.; England, L.; Rao, K.; Deane, J.; Collins, J. The FFC Cambridge Process for Production of Low Cost Titanium and Titanium Powders. Key Eng. Mater. 2010, 436, 75–83. [Google Scholar] [CrossRef] [Green Version]
  9. Fray, D.J.; Jiao, S. Treatment of Titanium Ores. Chinuka Limited. GB 2472496. 28 July 2010. Available online: https://www.mysciencework.com/patent/download/treatment-titanium-ores-EP2462251B1/EP2462251B1 (accessed on 4 December 2021).
  10. Ning, X.; Xiao, J.; Jiao, S.; Zhu, H. Anodic dissolution of titanium oxycarbide TiCxO1−x with different O/C ratio. J. Electrochem. Soc. 2019, 166, E22–E28. [Google Scholar] [CrossRef]
  11. Popov, B.N.; Kimbl e, M.C.; White, R.E.; Wendt, H. Electrochemical behaviour of titanium(II) and titaniun(III) compounds in molten lithium chloride/potassium chloride eutectic melts. J. Appl. Electrochem. 1991, 21, 351–357. [Google Scholar] [CrossRef]
  12. Zhu, H. Rare Earth Metal Production by Molten Salt Electrolysis. In Encyclopedia of Applied Electrochemistry; Springer: New York, NY, USA, 2014; pp. 1765–1772. [Google Scholar]
  13. Head, R.B. Electrolytic Production of Sintered Titanium from Titanium Tetrachloride at a Contact Cathode. J. Electrochem. Soc. 1961, 108, 806. [Google Scholar] [CrossRef]
  14. Haarberg, G.M.; Rolland, W.; Sterten, A.; Thonstad, J. Electrodeposition of titanium from chloride melts. J. Appl. Electrochem. 1993, 23, 217–224. [Google Scholar] [CrossRef]
  15. Song, J.; Wang, Q.; Zhu, X.; Hou, J.; Jiao, S.; Zhu, H. The Influence of Fluoride Anion on the Equilibrium between Titanium Ions and Electrodeposition of Titanium in Molten Fluoride–Chloride Salt. Mater. Trans. 2014, 55, 1299–1303. [Google Scholar] [CrossRef] [Green Version]
  16. Norikawa, Y.; Yasuda, K.; Nohira, T. Electrodeposition of Titanium in a Water-Soluble KF–KCl Molten Salt. Mater. Trans. 2017, 58, 390–394. [Google Scholar] [CrossRef] [Green Version]
  17. Réjasse, F. Study of the Reactivity of Group Ivb Metallic Dioxides in the Presence of Carbon by a (Micro)-Structural Approach: Application to the Thermodynamic Modelling of Ternary Phase Diag. Ph.D. Thesis, University of Limoges, Limoges, France, 2 December 2015; pp. 125–243. [Google Scholar]
  18. Gendre, M.; Maître, A.; Trolliard, G. Synthesis of zirconium oxycarbide (ZrCxOy) powders: Influence of stoichiometry on densification kinetics during spark plasma sintering and on mechanical properties. J. Eur. Ceram. Soc. 2011, 31, 2377–2385. [Google Scholar] [CrossRef]
  19. Geran, S.; Chamelot, P.; Serp, J.; Gibilaro, M.; Massot, L. Electrochemistry of Uranium in Molten LiCl-LiF; Elsevier: Amsterdam, The Netherlands, 23 April 2020. [Google Scholar]
  20. Ramaley, L.; Krause, M.S. Theory of square wave voltammetry. Anal. Chem. 1969, 41, 1362–1365. [Google Scholar] [CrossRef]
  21. Dring, K. Challenges and Opportunities in Titanium Metal Production, Is There Still Any Interest/Need to Replace Kroll Production? Norsk Titanium AS: Oslo, Norway; pp. 12–14. Available online: https://www.okabe.iis.u-tokyo.ac.jp/workshop/090219_Workshop_SanFrancisco_for_Okabe_lab_HP/parts/doc.C.pdf (accessed on 4 November 2021).
Figure 1. The diagram summarizes the process of synthesis and shaping of the titanium oxycarbide anode.
Figure 1. The diagram summarizes the process of synthesis and shaping of the titanium oxycarbide anode.
Materproc 05 00063 g001
Figure 2. Cyclic voltammograms on a tungsten electrode (S = 0.2 cm2) in LiCl-LiF (black) and LiCl-LiF-TiF3 (0.62 mol/kg) (blue) and anodic polarization curve of TiCxOy (Red), at 853 K with the scan rate of 100 mV·s−1.
Figure 2. Cyclic voltammograms on a tungsten electrode (S = 0.2 cm2) in LiCl-LiF (black) and LiCl-LiF-TiF3 (0.62 mol/kg) (blue) and anodic polarization curve of TiCxOy (Red), at 853 K with the scan rate of 100 mV·s−1.
Materproc 05 00063 g002
Figure 3. Square wave voltammograms on a tungsten electrode (0.12 cm2) in LiCl-LiF-TiF3 (7.25 × 10−2 mol·kg−1) at 853 K under various frequencies. Insert linear relationship of Ti3+ peak current density versus square root of frequency.
Figure 3. Square wave voltammograms on a tungsten electrode (0.12 cm2) in LiCl-LiF-TiF3 (7.25 × 10−2 mol·kg−1) at 853 K under various frequencies. Insert linear relationship of Ti3+ peak current density versus square root of frequency.
Materproc 05 00063 g003
Figure 4. X-ray diffraction of titanium metal deposited on the cathode under the current density of 60 mA·cm−2 in LiCl-LiF-TiF3 (0.62 mol/kg) at 853 K (580 °C). Insert image of titanium deposit.
Figure 4. X-ray diffraction of titanium metal deposited on the cathode under the current density of 60 mA·cm−2 in LiCl-LiF-TiF3 (0.62 mol/kg) at 853 K (580 °C). Insert image of titanium deposit.
Materproc 05 00063 g004
Figure 5. SEM micro-morphology of titanium metal deposit on the cathode under the current deposit of 60 mA·cm−2 in LiCl-LiF-TiF3 (0.62 mol/kg) at 853 K (580 °C).
Figure 5. SEM micro-morphology of titanium metal deposit on the cathode under the current deposit of 60 mA·cm−2 in LiCl-LiF-TiF3 (0.62 mol/kg) at 853 K (580 °C).
Materproc 05 00063 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Malek, B.; Serp, J.; Doreau, F.; Miguirditchian, M.; Vandenhende, M.; Pradeilles, N.; Lepetitcorps, Y.; Maitre, A. Production of Metallic Titanium by Electrowinning in Molten Salts of Titanium Oxycarbide Anode. Mater. Proc. 2021, 5, 63. https://doi.org/10.3390/materproc2021005063

AMA Style

Malek B, Serp J, Doreau F, Miguirditchian M, Vandenhende M, Pradeilles N, Lepetitcorps Y, Maitre A. Production of Metallic Titanium by Electrowinning in Molten Salts of Titanium Oxycarbide Anode. Materials Proceedings. 2021; 5(1):63. https://doi.org/10.3390/materproc2021005063

Chicago/Turabian Style

Malek, Btissem, Jerome Serp, Franck Doreau, Manuel Miguirditchian, Marion Vandenhende, Nicolas Pradeilles, Yann Lepetitcorps, and Alexandre Maitre. 2021. "Production of Metallic Titanium by Electrowinning in Molten Salts of Titanium Oxycarbide Anode" Materials Proceedings 5, no. 1: 63. https://doi.org/10.3390/materproc2021005063

Article Metrics

Back to TopTop