Qiwen Zhangabcd,
Yong Fan*abcd,
Yimin Zhangabcd,
Peng Liuabcd,
Hong Liuabcd,
Zihanyu Zhangabcd and
Xiuqiong Fuabcd
aSchool of Resource and Environmental Engineering, Wuhan University of Science and Technology, Wuhan 430081, Hubei Province, China. E-mail: fanyong@wust.edu.cn
bState Environmental Protection Key Laboratory of Mineral Metallurgical Resources Utilization and Pollution Control, Wuhan University of Science and Technology, Wuhan 430081, Hubei Province, China
cCollaborative Innovation Center of Strategic Vanadium Resources Utilization, Wuhan University of Science and Technology, Wuhan 430081, Hubei Province, China
dHubei Provincial Engineering Technology Research Center of High Efficient Cleaning Utilization for Shale Vanadium Resource, Wuhan University of Science and Technology, Wuhan 430081, Hubei Province, China
First published on 22nd September 2025
Molten salt electrolysis is a promising method for producing metallic vanadium, overcoming the limitations of conventional processes, such as high purity requirements for raw materials and high impurity content. This study introduces an innovative approach utilizing soluble vanadium–aluminum alloy anodes in KCl–LiCl molten salt, where vanadium is dissolved as V2+ and V3+. Optimal conditions were identified as 0.3 A cm−2 current density, 500 °C temperature, and 6 h duration. High purity metallic vanadium was prepared under these conditions, and the electrolysis efficiency reached 89.53%. Electrochemical analysis revealed three reduction processes: Al(III) → Al, V(III) → V(II) and V(II) → V. In molten salts, vanadium ions existed in coordination forms as VCl2, VCl3, VCl42−, and VCl63−, while aluminum ions were present as AlCl4−. The process was confirmed as diffusion-controlled, with calculated V(III) diffusion coefficients. Nucleation mechanism analysis demonstrated dual vanadium deposition pathways: instantaneous and progressive nucleation. Fluoride ions were found to enhance reaction kinetics through chloride substitution in aluminum complexes, increasing free chloride availability for vanadium coordination and improving ion mobility. The developed method offers significant advantages in energy efficiency and product quality compared to traditional metallurgical approaches, providing insights for optimization of molten salt electrolysis systems in refractory metal production.
Among existing vanadium production systems, molten salt electrolysis technology is the most promising method.16–18 It has the advantages of low reaction temperature, high efficiency, and high product purity and is widely used for metal purification. In the traditional vanadium refining process, AlV85 alloy requires vacuum electron beam melting, which necessitates temperatures above 1500 °C. In contrast, the molten salt electrolysis process typically operates below 1000 °C, significantly reducing energy consumption. The molten salt electrolysis process is influenced by factors such as electrode properties, salt composition, and temperature. The choice of electrode material for electrolysis plays a significant role in the process.19–21 Methods like Fray–Farthing–Chen (FFC),22,23 OS,24 and Solid Oxygen Ionic Membrane (SOM) have been considered promising for producing vanadium.25 Recent research has focused on optimizing these methods, for example, by developing novel cathode structures for FFC, investigating different anode compositions and feeding mechanisms for OS, and exploring stable oxygen-ion-conducting membranes for SOM. However, the graphite anode is easily consumed during electrolysis. And vanadium oxides are almost insoluble in the molten chlorides. The solid electro-deoxidation process is very slow and vanadium oxides must be obtained by a complex process.
Vanadium–aluminum (V–Al) alloy, as a type of vanadium alloy, can be advantageous compared to traditional reduction processes, which use vanadium oxides. During the preparation of V–Al alloy, by-products of aluminum can be easily washed away by acid or alkali. And the reduction process using aluminum lowers the reaction energy consumption, thus improving the yield of the metallic vanadium. However, the use of V–Al alloy also has certain limitations.26 During the electrolysis process, impurities such as Al may deposit on the cathode, which may lead to a decrease in the purity of the obtained metallic vanadium. To address this issue, a viable strategy involves the introduction of potassium fluoride (KF) into the molten salt system. Under certain conditions, KF demonstrates the capability to form stable coordination complexes with aluminum ions (Al3+), thereby modifying the coordination architecture of the molten salt medium.27 This structural reorganization effectively inhibits the electrochemical deposition of metallic aluminum through two primary mechanisms.
The molten salt electrolyte, as a carrier for the electrolysis process, also significantly affects the reaction. Alkali metal and alkaline earth metal chlorides are the most stable and inexpensive molten salts and are ideal for the molten salt system. Due to the relatively low decomposition voltage of CaCl2,28 the electrochemical behavior in CaCl2-based29 electrolytes is accompanied by calcium deposition, which interferes with the electrolysis process. Compared to the NaCl–KCl30,31 electrolyte, the KCl–LiCl32 system offers a lower eutectic temperature. Owing to the small ionic radius and high migration rate of Li ions, it exhibits good ionic conductivity, thereby reducing energy consumption costs to some extent. Additionally, the KCl–LiCl33 system possesses a wide electrochemical window. The standard electrode potentials of K and Li are lower than that of vanadium, meaning their deposition does not interfere with the reaction. To summarize, this study employs the KCl–LiCl system as the molten salt for the electrolysis process.
Therefore, this study proposes a simple, low-energy method for the production of metallic vanadium through electrolysis of soluble V–Al alloy anodes in a KCl–LiCl molten salt system. The schematic diagram is shown in Fig. 1. The study optimizes electrolysis conditions such as anode current density, reaction temperature, and time, and investigates the reduction and diffusion processes in the KCl–LiCl molten salt through various electrochemical tests. The diffusion coefficients and nucleation patterns of vanadium ions are calculated. The addition of fluoride (F) ions changes the electrochemical properties and reduction process, and their effects on the product are analyzed. Additionally, molecular dynamics (MD) and density functional theory (DFT) are combined to study the bonding processes in the reaction system through radial distribution functions and analyze the atomic interactions between ions during the electrolysis process.
![]() | ||
Fig. 2 (a) XRD of V–Al alloy prepared from sodium vanadate; (b) SEM of V–Al alloy prepared from sodium vanadate and the results of EDS analysis at point A. |
Element | V | Al | Fe | C | N | O |
---|---|---|---|---|---|---|
Composition (wt%) | 66.65 | 33.14 | 0.08 | 0.07 | 0.02 | 0.12 |
![]() | (1) |
mc = (m0 − m′) × αV | (2) |
Under the conditions of an anode current density of 0.3 A cm−2 and a cathode current density of 0.2 A cm−2, electrolysis of the V–Al alloy was performed, and the cell voltage and electrolysis time were recorded, as shown in Fig. 4a. It can be observed that the electrolysis process can be divided into three stages. In the initial stage, the voltage rises continuously, which may be due to the presence of some moisture in the molten salt. In the second stage, as there is no molten salt, the voltage remains at a high level for a period of time. During this process, vanadium is almost not electro-deposited. This means that there is an induction period (IP) for the electro-deposition of vanadium.34 In the third stage, the current rapidly decreases and then stabilizes, indicating that due to the electrochemical dissolution of V–Al, vanadium ions are introduced into the molten salt. The concentration of vanadium ions increases rapidly until it reaches saturation.
The anode current density determines the dissolution rate of the V–Al soluble anode. As shown in Fig. 4b, the influence of anode current densities ranging from 0.1 A cm−2 to 0.4 A cm−2 on metallic vanadium production was studied. A constant current electrolysis was performed at a fixed cathode current density of 0.2 A cm−2. With the increase in anode current density, the IP time significantly decreased. This is due to the rapid electrochemical dissolution of the V–Al soluble anode at higher anode current densities. As shown in Fig. 4c, when the anode current density increased from 0.1 A cm−2 to 0.4 A cm−2, the mass of metallic vanadium also increased accordingly, from 0.23 g to 0.34 g. Meanwhile, the electrolytic efficiency firstly increased and then decreased, which may be due to the lower concentration of vanadium ions in the molten salt at lower anode current densities, thereby increasing the IP for vanadium electro-deposition and reducing the electrolytic efficiency. When the anode current density is too high, the concentration of vanadium ions rapidly increases, and the ions cannot be completely deposited on the electrode, resulting in some vanadium loss. Therefore, the optimal anode current density is 0.3 A cm−2.
Reaction temperature is also an important factor affecting electrolysis efficiency. As shown in Fig. 4d, the electrolytic efficiency increased from 77.74% at 450 °C to 87.03% at 500 °C. However, when the temperature increased further to 550 °C, the electrolytic efficiency significantly decreased to 85.59%. This is because, at higher temperatures, vanadium ions in the molten salt will volatilize along with chloride ions, thus reducing the actual concentration of vanadium ions and lowering the electrolytic efficiency. As shown in Fig. 4e, metallic vanadium can be obtained in a certain amount during the electrolysis process from 4 h to 8 h. This indicates that metallic vanadium can be deposited after 4 h.
Based on the above results, the optimal anode current density and electrolysis temperature are 0.3 A cm−2 and 500 °C, respectively. The electrolysis time is selected 6 h. The obtained product was analyzed by SEM and EDS (Fig. 4f), the metallic vanadium prepared using this method has a high purity. ICP detection was performed on it, and the impurity content of the product was found to be less than 0.11%.
To further verify the above conclusion, square wave voltammetry (SWV) was performed (Fig. 5c), with a scanning frequency of 30 Hz. Clearly, three reduction peaks, R1, R2, and R3, can be observed, indicating that three reduction reactions occurred, and there is a good correspondence with the CV curve. The number of electrons transferred corresponding to the reduction peaks can be calculated using eqn (3):35,36
![]() | (3) |
The calculation of the SWV curve after Gaussian fitting yields n1 = 2.63 ≈ 3, n2 = 1.15 ≈ 1, and n3 = 1.71 ≈ 2. Therefore, R1 transfers three electrons, R2 transfers one electron, R3 transfers two electrons. It can be preliminarily determined that R1 is the process of Al(III) → Al, R2 is the process of V(III) → V(II), R3 is the process of V(II) → V.
The most commonly used electrolytes in electrolysis are chlorides and fluorides. Compared to Cl−, F− exhibits stronger coordination capability. Adding F− to chlorides triggers ligand substitution, altering the coordination structure of vanadium ions in the molten salt. This modification consequently affects the electrochemical properties of metal ions in the molten salt and their reduction processes. To investigate the electrochemical behavior of the molten salt after the addition of F ions, 10 g of KF was added to 100 g of KCl–LiCl molten salt, and the electrolyte was subjected to electrolysis for 6 h at 500 °C. The CV curve obtained is shown in Fig. 5d. It can be observed that after the addition of F ions, the potential of R1, R2 and R3 shifted leftward. The potential difference between R1 and R2 increased from 0.26 V to 0.4 V, which can better separate vanadium and aluminum. In addition, a new reduction peak R′ was formed in the KCl–LiCl–KF molten salt, suggesting the formation of VCliF6−i3−,37 while the bonding between Al and F was included in R1, which was not obvious.
Meanwhile, we conducted MD simulations of the molten salt electrolysis process for vanadium production using the Forcete module of Materials Studio (MS). The study combines the Universal force field with the charge equilibration method, utilizing electrostatic charge calculations within a constant particle number volume temperature (NVT) ensemble. Employing the Andersen thermostat with a time step of 1.0 ps, the total simulation duration was set to 3000 ps. Molecular dynamics trajectories were sampled at 0 ps, 1000 ps, 2000 ps, and 3000 ps for subsequent analysis. As shown in the Fig. 6d, the structure of KCl–LiCl after 3000 ps of MD simulation can be observed that vanadium ions and aluminum ions form coordination structures with chloride (Cl) ions. The form of vanadium ions are VCl63−, VCl42−, VCl2, VCl3. The form of Al ions are AlCl4−. That is consistent with the results of the Raman analysis mentioned above. And Fig. 6e shows that the energy fluctuation in the simulated system is minimal, indicating that the system is in a state of equilibrium throughout the simulation process.
Density of States (DOS) analysis based on band theory helps further analyze the interactions in the molten salt. When the density of states of different orbitals near the valence band superimposes and peaks at the same energy, it is considered that there is interaction between ions, which is a clear sign of bonding, and this peak is called a hybridization peak.40 Therefore, we added F ions to analyze the interactions between ions in the molten salt. As shown in Fig. 6f, the electronic orbitals near the lowest energy of −19 eV are occupied by F-2s orbitals, while the orbitals at an intermediate energy of −12 eV are occupied by Cl-2s orbitals. Near the Fermi level, the orbitals are mainly occupied by Cl-2p orbitals, with some hybridization with F-2p, V-3d, Al-3s, and Al-3p orbitals. This indicates that there are interactions between Cl, F, V, and Al in the molten salt.
The coordination tendencies of different metal cations with F− vary. To investigate the influence of F− on cation bonding in the KCl–LiCl system and reveal the structural evolution patterns and formation mechanisms of various ions in the KCl–LiCl–KF system, this study employed radial distribution function analysis to determine the probability of atomic occurrence within specific bonding regions, thereby identifying the bonding sequence of distinct ions. Furthermore, existing studies show that peaks below 3.5 Å are mainly formed by chemical bonds, while peaks above 3.5 Å are primarily formed by Coulomb and van der Waals interactions.41 As shown in Fig. 6g, V3+ forms a strong peak at 2.37 Å, and Al3+ forms a strong peak at 2.28 Å, indicating that V and Al form chemical bonds with Cl, with the V–Cl bond having a stronger atomic interaction compared to the Al–Cl bond, making the V–Cl bond more likely to form. As shown in Fig. 6h, Al3+ forms a strong peak at 2.43 Å with F, and the atomic interaction strength is higher than that of V–Cl and Al–Cl, suggesting that Al forms a chemical bond with F first, which facilitates the binding of more Cl ions to vanadium ions and promotes the dissolution of vanadium ions. Due to the high electronegativity of fluoride ions, their bond length with Al ions is shorter than that of V–Cl bonds, resulting in stable bond energy and Al ions are not easily deposited.
Afterwards, exploring the dynamic dissociation of structures through the breaking and reorganization of chemical bonds, simulation methods can accurately track the movement of each atom, facilitating the study of the evolution of chemical bonds.42 Fig. 6i presents snapshots of the molten salt structure at different time intervals. In the molten salt system, Li and K is randomly distributed in the simulation box, its small size and strong ionic nature prevent it from forming ion clusters with other ions. As time progresses, Cl ions in the molten salt gradually bond with vanadium ions, while aluminum ions directly bond with four Cl ions to form AlCl4−. In contrast, Fig. 6j shows that after the addition of F ions, aluminum ions preferentially bond with F ions, forming AlF4−, AlF52−, and AlF63− complexes. This process causes the Cl ions originally bonded to aluminum ions to coordinate with vanadium ions, thereby releasing more vanadium complexes, which aligns with the previously described results.
To further confirm the inhibitory effect of F− on Al ions, we conducted SEM/EDS analysis on the deposition products obtained before and after the addition of F−. The results are as follows:
A comparison between Fig. 7a and b reveals that in the absence of F−, the surface morphology exhibits distinct features characteristic of aluminum distribution. In contrast, after the addition of F−, the correlation with aluminum is significantly reduced. Further EDS analysis indicates a notable decrease in the relative percentage of aluminum in the deposits, accompanied by an increase in the relative content of vanadium. This compositional change is highly consistent with the observed shift in SEM morphological correlations. These results directly demonstrate that the introduction of F− effectively suppresses the aluminum deposition process.
![]() | ||
Fig. 7 SEM/EDS analysis results of cathode deposition products (a) before the addition of F−; (b) after the addition of F−. |
There are two methods for calculating the diffusion coefficient; when both the reactant and the product are soluble, eqn (4) is used for the calculation:43
![]() | (4) |
![]() | (5) |
The diffusion coefficient of V3+in the molten salt was calculated using eqn (4), and it was found to be 3.40 × 10−5 cm2 s−1.
Similarly, CV curves of molten salts after adding F ions were detected at different scan rates using the same method, as shown in Fig. 8c. The redox peak potential slightly shifted left with increasing scan rate, indicating that the corresponding reduction process is reversible. The relationship between the maximum current density of the reduction peaks and the square root of the scan rate was also investigated, as shown in Fig. 8d. After fitting the data, a linear relationship was observed, which indicates that the reduction reaction is still diffusion-controlled. The diffusion coefficient of V3+in the molten salt was calculated using eqn (4), and the diffusion coefficient of V3+ in the molten KCl–LiCl–KF salt was found to be 4.08 × 10−5 cm2 s−1.
To further confirm the above findings, SWV tests were performed at different frequencies. The results are shown in Fig. 8e. As the scan frequency increases, the peak potential gradually shifts negatively, indicating that the oxidation-reduction reaction is reversible. Moreover, as shown in Fig. 8f, the peak current density exhibits a linear relationship with the square root of the frequency. This is consistent with the results from CV. The diffusion coefficient of V3+ ions can also be calculated using eqn (6):44
![]() | (6) |
The diffusion coefficient of V3+ in the molten KCl–LiCl molten salt was calculated to be 2.90 × 10−5 cm2 s−1.
Similarly, in the molten salt after adding F ions, SWV tests were performed at different scan frequencies, as shown in Fig. 8g. The number of reduction peaks is consistent with the CV curve, and the reduction peak R1 is clearly visible. Moreover, as shown in Fig. 8h, the peak current density exhibits a linear relationship with the square root of the frequency. Thus, using eqn (6), the diffusion coefficient of V3+ ions in the molten KCl–LiCl–KF salt was calculated to be 3.45 × 10−5 cm2 s−1.
Furthermore, in MD simulations, the slope of the MSD curve, divided by 1/6, corresponds to the diffusion coefficient fitted according to Einstein diffusion equation, as shown in eqn (7):45
![]() | (7) |
The MSD curve of the simulation system and the fitting curve are shown in Fig. 8i and j. The diffusion coefficient of unfluorinated V3+ is calculated to be 3.36 × 10−5 cm2 s−1, while the diffusion coefficient of fluorinated V3+ is 7.44 × 10−5 cm2 s−1, which is on the same order of magnitude as the electrochemical tests, and the diffusion rate increases after the introduction of F.
Based on the above data, we analyzed the diffusion coefficients of three methods in Fig. 8k. It can be concluded that after adding F ions, the diffusion rate of V3+ increased overall. This is due to the formation of a new reduction peak, which increases the concentration of vanadium ions and accelerates the mass transfer rate.
Generally, there are two different types of nucleation: instantaneous nucleation and continuous nucleation. The nucleation process model is as follows:46,47
For instantaneous nucleation:
![]() | (8) |
For continuous nucleation:
![]() | (9) |
Calculating the theoretical model and experimental values in the dimensionless plots of (I/Im)2 and (t/tm) using eqn (8) and (9). It can be concluded that the nucleation of vanadium ions involves both instantaneous nucleation and continuous nucleation by fitting and analyzing the data (Fig. 9b and e). The obtained metal vanadium product was washed with deionized water, soaked in dilute hydrochloric acid to remove aluminum and other impurities, and then vacuum-dried and collected. The product was characterized by scanning electron microscopy (SEM), as shown in Fig. 9c and f. The product exhibited two morphologies: dendritic and foamy, where the dendritic morphology was formed by continuous nucleation and the foamy morphology by instantaneous nucleation.
The formation of the aforementioned morphologies is primarily related to the current density and vanadium ion concentration during the electrolysis process. In the initial stages of electrolysis, the high ion concentration prevents ion depletion near the cathode. At this point, the rate of crystal growth outpaces the rate of nucleation during deposition, favoring rapid nucleation and the development of simpler morphologies, thereby resulting in a foamy structure. As electrolysis progresses and ion concentrations decrease, concentration polarization occurs near the cathode. Under these conditions, the nucleation of vanadium ion deposition becomes more challenging, and crystal growth is hindered, leading to the formation of dendritic structures.
Finally, the mechanism of the whole electrolysis process is analyzed in Fig. 10. During the anodic dissolution process, an external voltage is applied to the anode by a DC power source, causing the V–Al alloy to undergo an oxidation reaction under electrochemical influence. Vanadium and aluminum lose electrons and enter the electrolyte, transforming into V2+, V3+, and Al3+. The oxidation process of vanadium is influenced by the electrolyte composition and current density. At lower current densities, the voltage is reduced, and the oxidation is predominantly dominated by V3+ ions. The dissolution process is as follows:
V → V2+ + 2e− | (10) |
V → V3+ + 3e− | (11) |
Al → Al3+ + 3e− | (12) |
Ultimately, the complexes in the molten salt gradually migrate toward the cathode under the influence of the electric current. The addition of an appropriate amount of F− accelerates ion mass transfer rates, thereby promoting the deposition of metal at the cathode. On the cathode surface, vanadium ions in the molten salt are reduced to metallic vanadium. The main reaction equation of vanadium complexes during cathodic deposition is as follows:
VCl63− + e− → VCl42− + 2Cl− | (13) |
VCl42− + 2e− → V + Cl− | (14) |
While aluminum ions largely coordinated with F− remain in ionic form under specific potential conditions, enabling the purified deposition of metallic vanadium. A small amount of deposited metallic aluminum undergoes thermal reduction reactions with undeposited vanadium ions at high temperatures, where V2+ or V3+ are reduced to V or V2+, and Al is oxidized back to Al3+, re-entering the electrolyte to continue participating in the reaction cycle. The main reaction equation is as follows:
Al3+ + 3e− → Al | (15) |
Al + 3V3+ → 3V2+ + Al3+ | (16) |
2Al + 3V2+ → 2Al3+ + 3V | (17) |
Subsequently, the diffusion coefficient and nucleation mode of V3+ were analyzed. All of the reversible processes are controlled by diffusion. The diffusion coefficient of V3+ was calculated by CV, SWV and MDS. Before adding F ions, the calculated diffusion coefficients were 3.40 × 10−5 cm2 s−1, 2.90 × 10−5 cm2 s−1 and 4.12 × 10−5 cm2 s−1. After adding F ions, the calculated diffusion coefficients were 4.08 × 10−5 cm2 s−1, 3.36 × 10−5 cm2 s−1 and 7.44 × 10−5 cm2 s−1. It can be concluded that with the addition of F ions, the diffusion rate of V3+ accelerates. And it is determined that vanadium ions undergo both instantaneous nucleation and continuous nucleation on the tungsten electrode surface. This work provides a new and feasible method for the production of metallic vanadium.
This journal is © The Royal Society of Chemistry 2025 |