Next Article in Journal
Chiral Perovskite Single Crystals: Toward Promising Design and Application
Previous Article in Journal
Particleboards with Various Biomass Residues
Previous Article in Special Issue
Enhanced Photoelectrocatalytic Performance of ZnO Nanowires for Green Hydrogen Production and Organic Pollutant Degradation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Preparation of Metallic Zr from ZrO2 via Carbothermal and Electrochemical Reduction in Molten Salts

1
Department of Radiochemistry, China Institute of Atomic Energy, Beijing 102413, China
2
Panxi Institute of Vanadium and Titanium Inspection and Testing, Panzhihua 617000, China
3
State Key Laboratory of Vanadium and Titanium Resources Comprehensive Utilization, Pangang Group Research Institute Co., Ltd., Panzhihua 617000, China
*
Author to whom correspondence should be addressed.
Materials 2025, 18(11), 2634; https://doi.org/10.3390/ma18112634
Submission received: 7 April 2025 / Revised: 21 May 2025 / Accepted: 26 May 2025 / Published: 4 June 2025

Abstract

:
Zirconium, a critical rare metal with exceptional corrosion resistance and nuclear applications, is conventionally produced via the energy-intensive Kroll process. The electrolysis of ZrCxOy soluble anodes has been extensively investigated due to its advantages in having a short process flow and resulting in high-quality products. In particular, during the electrolysis of zirconium oxycarbide with a C:O molar ratio of 1:1, gaseous CO can be released, and no residual anodes are generated, which is extremely appealing. In this regard, this paper explores the feasibility of preparing zirconium metal through high-temperature vacuum reduction to produce zirconium oxycarbide using ZrO2 as the raw material, followed by direct molten-salt electrolysis. Firstly, the reduction products were characterized using an X-ray diffractometer (XRD) and a scanning electron microscope (SEM). The results showed that under a vacuum of <10 Pa at 1750 °C, the reduction products mainly consisted of ZrCxOy and a small amount of ZrO2, and they exhibited good electrical conductivity (0.0169 Ω·cm). Subsequently, the cyclic voltammetry test results of the reduction products revealed the reversible redox behavior of ZrCxOy. There were characteristic oxidation peaks at −0.53 V and −0.01 V (vs. Pt), corresponding to the formation of Zr2+ and Zr4+, respectively, and a reduction peak at −1.51 V, indicating the conversion from Zr2+ to Zr. Finally, β-zirconium metal with a purity of 99.2 ± 0.3 wt.% was obtained through potentiostatic electrolysis, and its quality met the R60704 grade specified in ASTM B551-12 (2021). This study offers a novel approach for the short-flow preparation of zirconium metal, which is conducive to expanding its applications.

1. Introduction

Zirconium is an important rare metal with excellent corrosion resistance, good mechanical properties, and a low thermal neutron absorption cross section; this material is widely used in the aerospace, nuclear, and chemical industries [1,2]. Global zirconium resources are abundant, but mainly in the form of zircon (ZrSiO4), and zircon is often symbiotic with other minerals, which makes it difficult to substantially increase zirconium extraction [3,4]. Currently, zirconium metal is mainly prepared industrially through the magnesium thermal reduction of ZrCl4 (Kroll process), but this process is time consuming, necessitates high energy consumption, and is expensive [5,6]. Therefore, compared with thermal reduction processes, molten salt electrolysis has received more attention because of its time efficiency and the good quality of the obtained products [7,8].
According to the different raw materials used in electrolysis, Zr metal preparation can be divided into zirconium halide and oxide electrolysis process; the former is mainly used in the preparation of zirconium metal via the direct molten salt electrolysis of halide raw materials, such as ZrCl4, and K2ZrF6, but it is still poorly efficient and expensive [9,10]; the latter mainly uses ZrO2 as the raw material to prepare zirconium metal via molten salt electrolysis. Li et al. prepared Zr–Mg alloys via the electrochemical deoxidization of ZrO2–MgO, using the FFC Cambridge process. Unfortunately, this process has also low efficiency, and the obtained products contain excessive oxygen impurities [11,12]. A new process for the electrolytic preparation of zirconium metal was proposed by Shang et al. [13]. They prepared a conductive ZrCxOy anode via electrochemical reduction using a ZrO2 and C mixture as the cathode and graphite as the anode; they then electrolyzed ZrCxOy to obtain zirconium metal, while a carbon steel rod was used as the cathode. The zirconium metal obtained via this process has high purity, and the electrolysis process is stable, but the preparation of ZrCxOy soluble anodes is affected by a low efficiency and poor material uniformity. Actually, ZrO2 can be reduced to ZrCxOy using ZrC and Zr above 1600 °C [14]. Li et al. systematically investigated the influence of different carbon-to-ZrO2 ratios on the reduction product ZrCxOy at 1600 °C under argon protection. Their findings indicated that when the molar ratio of ZrO2 to C is 1:2, the reduction products are ZrCxOy and ZrO2, rather than the target product ZrC0.5O0.5. When the molar ratio of ZrO2 to C is 1:2.6, a pure ZrC0.79O0.21 phase can be obtained. Using this as the raw material, they successfully prepared zirconium metal via electrolysis [15]. However, a certain amount of residual carbon is generated at the anode, which can reduce electrolysis efficiency and contaminate the product quality as the electrolysis progresses. Jiao et al. demonstrated that when the carbon-to-oxygen molar ratio in titanium oxycarbide (TiCxO1−x) is 1:1 (x = 0.5), electrolysis with a soluble anode eliminates residual anode formation and enables metallic titanium deposition on the cathode [16]. Inspired by this, it is hypothesized that ZrC0.5O0.5 electrolysis products may also produce minimal residual anodes. Nevertheless, it remains unclear whether ZrO2 can be directly carbothermally reduced to ZrC0.5O0.5 at higher temperatures and under vacuum conditions. Additionally, it is unknown whether the reduction products can be directly used in molten-salt electrolysis to prepare zirconium metal.
This study proposes a carbothermal and electrochemical reduction process for the preparation of zirconium metal, focusing on the electrochemical dissolution and deposition of ZrCxOy prepared through the carbothermal reduction of ZrO2 and the ZrCxOy soluble anode. The results demonstrate that ZrO2 can be reduced to conductive ZrCxOy through reaction with graphite at high temperatures, and a metallic Zr was electrolyzed from this ZrCxOy as an anode in molten salts.

2. Materials and Methods

2.1. Carbothermal Reduction of ZrO2

To gain in-depth insights into the reduction process of ZrO2 by carbothermal reduction, the thermodynamic parameter conditions for the reduction of zirconium oxycarbide were calculated. Although direct thermodynamic data for ZrC0.5O0.5 are lacking, the formation of an infinite solid solution between ZrO and ZrC enables us to deduce the synthesis conditions of ZrC0.5O0.5 by analyzing the reduction of ZrO2 to these two phases [14,17]. By using HSC Chemistry 6.0 software, thermodynamic calculations for the reduction of ZrO2 resulted in the following reaction:
2ZrO2 + 4C→ZrC + ZrO + 3CO(g)       ΔGθ = 1518.1 − 0.689T
Over the 0–2000 °C range, the positive standard Gibbs free energy of Reaction 1 indicates non-spontaneity under ambient pressure. Under a vacuum, the Gibbs free energy is modified to account for CO partial pressure [18]:
ΔG = ΔGθ + RT ln(PCO/P0)3 = 1518.1 − 0.689T + 0.02492T ln(PCO/P0)
Figure 1 illustrates the temperature-dependent Gibbs free energy at varying vacuum levels. Increasing temperature lowers ΔGθ, with vacuum enhancement (i.e., reduced total pressure) accelerating this decline. At 0.1 Pa, spontaneity is achieved at 1500 °C. Balancing furnace operational limits, the optimized reduction parameters were set as 1750 °C, 5 h, and a vacuum level <10 Pa.
Under the aforementioned reduction conditions, specific amounts of ZrO2 powder (AR, Shanghai Macklin Biochemical Technology Co., Shanghai, China) and graphite powder (<45 µm, Qingdao Tianheda graphite Co., Ltd., Qingdao, China) were first weighed and placed in a three-dimensional mixer (SYH-5, Changzhou Yineng drying equipment Co., Ltd., Changzhou, China) with ZrO2:C molar ratio of 1:2. After 4 h of mixing, 3 wt.% polyvinyl alcohol binder was added, and mixing was continued for 2 h. Then, about 50 g of the mixture was placed in a 30 mm diameter grinding tool under a four-column hydraulic press (Y32-250T, Tengzhou Hairun Machine Tool Co., Ltd., Tengzhou, China); a pressure of 20 MPa was applied for 20 s to obtain the required block. This column was then placed in a 120 °C oven for drying for 12 h and later transferred to a vacuum furnace (JVIM, Shenyang Jinyan New Material Preparation Technology Co., Ltd., Shenyang, China) for carbothermal reduction. After the reduction, the reduction products were taken out of the furnace for weighing and were characterized using X-ray diffraction (XRD, X’pert PRO, PANalytical, Almelo, The Netherlands) and scanning electron microscopy in combination with an energy-dispersive spectrometer (SEM-EDS, SIGMA 500, Carl Zeiss AG, Jena, Germany).

2.2. Electrochemical Measurements and Electrodeposition

The electrochemical behavior of the reduced products was characterized by cyclic voltammetry (CV) in NaCl-KCl (equimolar ratio) molten salt under the protection of argon gas using an electrochemical workstation (PARSTAT 4000A, AMETEK, Inc., Berwyn, IL, USA) connected to a three-electrode system. ZrCxOy was embedded in a small hole of a molybdenum rod with a diameter of 4 mm (hole diameter of 2 mm), which was used as the working electrode. A platinum rod with a diameter of 4 mm and a graphite rod with a diameter of 4 mm were used as the quasi-reference and counter electrodes, respectively.
After the electrochemical measurements, the hollow reduction product was connected to a graphite rod, which was then connected to the positive terminal of a power amplifier (BOP 20-20M, KEPCO, Inc., New York, NY, USA), while the negative terminal was connected to a molybdenum rod with a diameter of 8 mm. Constant-potential electrolysis was performed at a reduction potential of −1.6 V (vs. Pt) to study the dissolution characteristics of ZrCxOy. After electrolysis, the cathode and anode products were removed from the molten salt and cooled to room temperature in argon; they were then washed in 0.5 wt.% hydrochloric acid to remove the electrolyte trapped within the cathode and anode products; the hydrochloric acid was subsequently washed off with deionized water to obtain the cathodic products and anodic residue. The cathode products were also characterized via XRD and SEM. In addition, inductively coupled plasma emission spectrometry (ICP-AES, NexION 300D, PerkinElmer, Waltham, MA, USA) was employed to analyze components such as Zr, Fe, and Cr in the deposited products [19,20]. For the determination of O, N, and H, an oxygen, nitrogen, and hydrogen analyzer (Leco ONH 836, Saint Joseph, MI, USA) was utilized.

3. Results and Discussion

3.1. Carbothermal Reduction

Figure 2a shows that the reduced product exhibits a black-colored, dense morphology. The main diffraction peaks of the reduced product are close to those of ZrC, but a local magnification of the (111) plane reveals that the diffraction peak of the reduced product is shifted by +0.01° compared to pure ZrC. In face-centered cubic ZrC, carbon atoms can be substituted by oxygen atoms to form ZrCₓOy solid solutions, which causes a positive shift in diffraction peaks—this phenomenon is consistent with the behavior of TiCₓOy, whose diffraction peaks lie between those of TiC and TiO [21]. Notably, the reduced product in Figure 2a still contains trace amounts of ZrO2. Given that low-valent zirconium oxides and ZrCₓOy with high oxygen content are unstable phases [22,23], and combining with the preferential formation of TiC during TiO2 reduction, the reduction pathway of ZrO2 is proposed as ZrO2→ZrC→ZrCxOy. Additionally, compared with the results of Li et al., the absence of diffraction peaks from residual carbon and the lower intensity of ZrO2 peaks in the reduced product—achieved via vacuum and higher-temperature reduction—indicate a more complete reduction reaction [15]. In addition, Figure 2b,c show that carbon (C) and oxygen (O) are generally uniformly distributed in zirconium oxycarbide, although localized regions with unreduced ZrO2 persist. The C/O molar ratios vary across the product: in Figure 2c, the ratio is 3.07 at position 1 and 1.47 at position 4, indicating local compositional heterogeneity. Notably, particles with higher C/O ratios exhibit finer sizes, while ZrO2 particles are coarser; regions adjacent to ZrO2 show lower C/O ratios, further supporting the reaction between ZrC and ZrO2 to form ZrCxOy. These findings suggest that ZrO2 is preferentially reduced to ZrC, which then reduces surrounding ZrO2 to ZrCxOy. However, the formation of large pores during reduction hinders the further conversion of ZrO2 to ZrCxOy, resulting in a C/O molar ratio >1 in zirconium oxycarbide.
Table 1 shows that the weight loss rate of the reduced products is 26.91%, and the reduction rate reaches 94.16% (while the theoretical weight loss rate is 28.58%). The resistivity of the reduction products (0.0169 Ω cm) is lower than that before reduction (0.0440 Ω cm), indicating that the reduction products have good electrical conductivity and are suitable for electrolysis as soluble anodes.

3.2. Electrochemical Behavior of Zirconium Oxycarbide

Figure 3 shows the CV curves obtained for the Mo and ZrCxOy electrodes in the KCl–NaCl melt at 750 °C. No other electrochemical signals are observed in the NaCl–KCl molten salts except for two pairs of redox peaks corresponding to Na+→Na and Mo2+→Mo at −2.2 V (vs. Pt) and 0.5 V (vs. Pt), respectively [21]. However, when using a ZrCxOy working electrode, two prominent oxidation peaks (O1 and O2) emerge at −0.20 V and −0.53 V (vs. Pt) during the anodic scan, accompanied by three reduction peaks (R1, R2, and R3) at −0.30 V, −0.78 V, and −1.51 V (vs. Pt) during the cathodic scan. According to Li et al., the oxidation of Zr3+ to Zr4+ occurs at approximately −0.20 V (vs. Pt) [15]; thus, O1/R1 can be assigned to the Zr3+/Zr4+ redox couple. Notably, the reduction potential of R1 is ~0.2 V higher than reported by Li et al. [15]. This positive shift arises from the rapid diffusion and dilution of dissolved Zr ions during the anodic scan, as the initial electrolyte contains no Zr species. Based on Shang et al., the potential difference between Zr4+/Zr3+ and Zr4+/Zr (or Zr+/Zr) is ~1.2 V [13]. Therefore, R3 is attributed to the reduction of Zr4+ (or Zr+) to metallic Zr. The absence of a corresponding oxidation peak for R3 is likely due to the large electrode area (4 mm diameter Mo rod embedded with ZrCxOy), which suppresses the reverse oxidation process. The potential difference between O1 and O2 is 0.38 V, significantly smaller than the expected 0.6 V for Zr2+→Zr4+ oxidation [15]. This suggests that O2 corresponds to the oxidation of ZrCxOy to Zr3+, with R2 as the complementary reduction peak. Considering that CO is the oxidation byproduct of ZrCxOy and the molar ratio of C to O is 1:1, as reported in reference [24], the electrode reactions during electrolysis are proposed as follows:
Anode: ZrC0.5O0.5 → Zr3+ + 0.5CO(g) + 3e
Cathode: Zr3+ + 3e → Zr
Overall reaction: ZrC0.5O0.5 →Zr + 0.5CO(g)
The above results indicate that no residual anode is generated when the C/O molar ratio in ZrCxOy is 1:1, whereas residual carbon forms when the C/O ratio exceeds 1:1. Combined with the findings in Figure 1, this residual carbon can react with unreduced ZrO2 via Reaction 6 to produce metallic zirconium and CO, suggesting that during the reduction process under a carbon-to-oxygen ratio of 1:1, ZrCxOy with a C/O molar ratio greater than 1:1 is formed. Although molten salt electrolysis under these conditions produces residual carbon, this residual carbon can further react with unreduced ZrO2 through electrolysis to form metallic zirconium and CO, thereby eliminating residual anode formation.
ZrO2 + 2C → Zr + 2CO(g)
Additionally, Figure 3b demonstrates that increasing the scan rate from 0.05 to 0.20 V s⁻¹ causes the oxidation potential of ZrCxOy to shift positively from −0.58 V to −0.53 V (vs. Pt), while the reduction potential shifts negatively. This behavior confirms the irreversible nature of the ZrCxOy dissolution and reduction processes [25,26].
The appearance of the cathode and anode products after electrolysis at −1.6 V (vs. Pt) for 3 h as well as the corresponding XRD patterns and SEM image of the cathodic products after washing are shown in Figure 4. The figure shows that the cathodic product wrapped by the electrolyte has been deposited on the Mo rod, while the anode has been wrapped by the electrolyte without undergoing dissolution. However, the residual anode after washing shows obvious dissolution traces and holes (Figure 4a), and the anode dissolution rate after washing and drying is 7.86%, which further demonstrates that ZrCxOy can be dissolved in the NaCl–KCl molten salts. The XRD results of the washed cathodic products indicate that the deposited product is pure β-phase metallic zirconium (Figure 4b), which is largely consistent with the findings of Li et al. [15]. The SEM–EDS results in Figure 4c show that the cathodic product consists of fine flakes and agglomerates of particulate matter, containing small amounts of impurity elements such as O, Cl, K, and Fe. The presence of Cl, K, and O might be attributed to hydrolysis during the washing process and residual electrolyte, while the Fe impurity could result from contamination by the cathode material.
The results presented in Table 2 indicate that the purity of the electrolytically obtained zirconium metal reaches 99.2 ± 0.3 wt.%. Impurity elements such as Fe + Cr, H, O, C, and N fully meet the requirements of the R60704 grade in the American standard ASTM B551-12 (2021) [27]. Considering that the theoretical molar ratio of C to O in ZrCxOy is 1:1, no residual carbon is generated during the electrolysis process. This results in a low carbon content in the products of the electrolysis. Although increasing the carbon-to-oxygen ratio can allow for the production of pure ZrCxOy, it also causes carbon contamination in the products. Therefore, maintaining the theoretical ratio enhances the quality of the zirconium metal. Furthermore, the continuous and stable operation of the electrolysis can be ensured by regularly removing undissolved ZrO2 from the anode surface with a scraper or other means. Overall, this powder can be directly used as a raw material for powder metallurgy, which can significantly reduce the cutting losses during the processing of zirconium-based materials. This circumvents the preparation of ZrCl4 and the use of expensive magnesium metal reductant in the production of zirconium by the Kroll process.
The above-mentioned research shows that ZrCxOy can be synthesized from ZrO2 through vacuum carbothermal reduction. Subsequently, zirconium metal products can be fabricated by directly conducting molten-salt electrolysis with ZrCxOy as a soluble anode. Compared with the direct electro-deoxidation of ZrO2 for zirconium metal production, this method is more efficient and yields higher-quality products, thus showing promising application prospects [11].

4. Conclusions

This study presents a novel carbothermal–electrolytic route for zirconium production, achieving direct conversion of ZrO2 to high-purity metal via a conductive oxycarbide intermediate. Key advancements include the following:
Thermodynamically optimized carbothermal reduction at 1750 °C/<10 Pa, yielding ZrCxOy with ideal electrical conductivity (0.0169 Ω·cm).
Electrochemical validation of ZrCxOy as a soluble anode, enabling reversible Zr2+/Zr4+ redox cycles and efficient metal deposition at −1.6 V (vs. Pt).
The production of high-purity zirconium, achieving a purity of 99.2 ± 0.3 wt.%, is characterized by the formation of fine flakes and agglomerates of particulate matter. This method circumvents the chlorination steps, thereby reducing the complexity of the production process.
This integrated approach paves the way for sustainable zirconium metallurgy, with potential for scaling to industrial capacities and extending to other refractory metals via oxycarbide intermediates.

Author Contributions

Conceptualization, W.S. and X.C.; methodology, W.S.; software, W.S.; validation, W.S., Y.J. and F.Z.; formal analysis, M.Y.; investigation, G.Y.; resources, X.C.; data curation, X.C.; writing—original draft preparation, W.S.; writing—review and editing, W.S.; visualization, X.C.; supervision, G.Y.; project administration, Y.J.; funding acquisition, G.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the China National Nuclear Corporation Leading Innovation Fund (Grant No. LC192209000803).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

Author Fuxing Zhu was employed by the company Pangang Group Research Institute Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Birch, R.M.; Douglas, J.O.; Britton, T.B. Characterization of local deformation around hydrides in Zircaloy-4 using conventional and high angular resolution electron backscatter diffraction. Mater. Charact. 2023, 202, 112988. [Google Scholar] [CrossRef]
  2. Xu, J.; Li, H.; Zhao, X.; Wu, J.; Zhao, B.; Zhao, H.; Wu, J.; Zhang, Y.; Liu, C. Zirconium based neutron absorption material with outstanding corrosion resistance and mechanical properties. J. Nucl. Mater. 2022, 567, 153763. [Google Scholar] [CrossRef]
  3. Perks, C.; Mudd, G. Titanium, zirconium resources and production: A state of the art literature review. Ore Geol. Rev. 2019, 107, 629–646. [Google Scholar] [CrossRef]
  4. Zeng, Y.; Liang, F.; Liu, J.; Zhang, J.; Zhang, H.; Zhang, S. Highly Efficient and Low-Temperature Preparation of Plate-Like ZrB2-SiC Powders by a Molten-Salt and Microwave-Modified Boro/Carbothermal Reduction Method. Materials 2018, 11, 1811. [Google Scholar] [CrossRef]
  5. Park, K.; Nersisyan, H.H.; Lee, H.; Hong, S.-I.; Cho, N.-C.; Lee, J.-H. Low-temperature synthesis of zirconium metal using ZrCl4–2Mg reactive mixtures. Int. J. Refract. Met. H 2012, 33, 33–37. [Google Scholar] [CrossRef]
  6. Li, S.; Che, Y.; Shu, Y.; He, J.; Song, J.; Yang, B. Review—Preparation of Zirconium Metal by Electrolysis. J. Electrochem. Soc. 2021, 168, 062508. [Google Scholar] [CrossRef]
  7. Li, M.; Liu, C.; Ting, A.; Xiao, C. A review on the extraction and recovery of critical metals using molten salt electrolysis. J. Environ. Chem. Eng. 2023, 11, 109746. [Google Scholar] [CrossRef]
  8. Xi, X.; Feng, M.; Zhang, L.; Nie, Z.-R. Applications of molten salt and progress of molten salt electrolysis in secondary metal resource recovery. Int. J. Miner. Metall. Mater. 2020, 27, 1599–1617. [Google Scholar] [CrossRef]
  9. Zhang, X.; Huang, J.; Min, D.; Wen, T.; Shi, Z.; Yang, S. Mechanism for electrochemical reduction of Zr(IV) in molten NaCl–KCl. Int. J. Hydrogen Energy 2022, 47, 6544–6551. [Google Scholar] [CrossRef]
  10. Li, S.; Che, Y.; Song, J.; Shu, Y.; He, J.; Xu, B.; Yang, B. Preparation of zirconium metal through electrolysis of zirconium oxycarbonitride anode. Sep. Purif. Technol. 2021, 274, 118803. [Google Scholar] [CrossRef]
  11. Li, N.; Peng, Y.; Chen, Z.; Xiong, W.; Sun, J.; Chen, Y.; Zhang, P.; Liu, M.; Li, S. Preparation of Mg-Zr alloys through direct electro-deoxidation of MgO-ZrO2 in CaCl2-NaCl molten salt. Electrochim. Acta 2021, 372, 137816. [Google Scholar] [CrossRef]
  12. Abdelkader, A.; Daher, A.; Abdelkareem, R.A.; El-Kashif, E. Preparation of Zirconium Metal by the Electrochemical Reduction of Zirconium Oxide. Metall. Mater. Trans. B 2007, 38, 35–44. [Google Scholar] [CrossRef]
  13. Shang, X.; Li, S.; Che, Y.; Shu, Y.; He, J.; Song, J. Novel extraction of Zr based on an in-situ preparation of ZrCxOy. Sep. Purif. Technol. 2021, 275, 118096. [Google Scholar] [CrossRef]
  14. Liu, H.; Man, Z.; Liu, J.; Wang, X.-G.; Zhang, G.-J. Solid solution and densification behavior of zirconium oxycarbide (ZrCxOy) ceramics via doping ZrO2 and Zr in ZrC. J. Alloys Compd. 2017, 729, 492–497. [Google Scholar] [CrossRef]
  15. Li, S.; Che, Y.; Song, J.; Shu, Y.; Xu, B.; He, J.; Yang, B. Electrolytic Preparation of Zirconium Metal from a Consumable Zirconium Oxycarbide Anode. Metall. Mater. Trans. B 2021, 52B, 3276–3287. [Google Scholar] [CrossRef]
  16. Jiao, S.; Zhu, H. Novel metallurgical process for titanium production. J. Mater. Res. 2006, 21, 2172–2175. [Google Scholar] [CrossRef]
  17. Katea, S.N.; Riekehr, L.; Westin, G. Synthesis of nano-phase ZrC by carbothermal reduction using a ZrO2–carbon nano-composite. J. Eur. Ceram. Soc. 2021, 41, 62–67. [Google Scholar] [CrossRef]
  18. Wang, J.; Jia, C.; Li, C.; Peng, X.-L.; Zhang, L.-H.; Liu, J.-Y. Thermodynamic Properties for Carbon Dioxide. ACS Omega 2019, 21, 19193–19198. [Google Scholar] [CrossRef]
  19. Hamm, D.; Ogle, K.; Olsson, C.-O.A.; Weber, S.; Landolt, D. Passivation of Fe–Cr alloys studied with ICP-AES and EQCM. Corros. Sci. 2002, 44, 1443–1456. [Google Scholar] [CrossRef]
  20. Mehdi, N.; Pouyan, S. A simple method for determining Hf in Zr and Zr alloys by ICP-AES. Nucl. Sci. Tech. 2009, 20, 340–343. [Google Scholar] [CrossRef]
  21. Long, W.; Gao, F.; Wang, D.; Zou, B.; Wang, X.; Wang, Y.; Zhan, F.; Wei, Y. Forming thermodynamics, structure, and electrical conductivity of TiCxOy compounds fabricated through the carbothermal reduction process. J. Alloys Compd. 2022, 892, 162201. [Google Scholar] [CrossRef]
  22. Pham, D.; Gai, F.; Rochester, J.; Dycus, J.H.; LeBeau, J.M.; Manga, V.R.; Corral, E.L. Thermodynamic assessment within the Zr–B–C–O quaternary system. J. Am. Ceram. Soc. 2023, 106, 3127–3140. [Google Scholar] [CrossRef]
  23. Efaw, C.M.; Vandegrift, J.L.; Reynolds, M.; McMurdie, S.; Jaques, B.J.; Hu, H.; Xiong, H.; Hurley, M.F. Characterization of zirconium oxides part I: Raman mapping and spectral feature analysis. Nucl. Mater. Energy 2019, 21, 100707. [Google Scholar] [CrossRef]
  24. Li, S.; Lv, Z.; He, J.; Song, J. Anodic dissolution of consumable ZrCxOy and cathodic deposition of zirconium in NaCl-KCl based melts. J. Mol. Liq. 2025, 420, 126831. [Google Scholar] [CrossRef]
  25. Zhang, K.; Xu, L.; Xia, D.; Meng, J.; Zhao, Z. A fundamental study on indium electrochemistry in molten NaCl-KCl eutectic: Toward new technology development for indium secondary resources recycling. Ionics 2023, 29, 4379–4387. [Google Scholar] [CrossRef]
  26. Wang, Z.; Tada, E.; Nishikata, A. In Situ Analysis of Scan Rate Effects on Pt Dissolution Under Potential Cycling Using a Channel Flow Double Electrode. Electrocatalysis 2015, 6, 179–184. [Google Scholar] [CrossRef]
  27. B551/B551M-12; Standard Specification for Zirconium and Zirconium Alloy Strip, Sheet, and Plate. ASTM International: West Conshohocken, PA, USA, 2021. [CrossRef]
Figure 1. Temperature-dependent Gibbs free energy at different vacuum levels.
Figure 1. Temperature-dependent Gibbs free energy at different vacuum levels.
Materials 18 02634 g001
Figure 2. XRD patterns (a), and SEM–EDS results (b,c) of the reduction products.
Figure 2. XRD patterns (a), and SEM–EDS results (b,c) of the reduction products.
Materials 18 02634 g002
Figure 3. CV curves obtained for the Mo electrode (a) and ZrCxOy electrode (b) in the KCl–NaCl melt at different scan rates.
Figure 3. CV curves obtained for the Mo electrode (a) and ZrCxOy electrode (b) in the KCl–NaCl melt at different scan rates.
Materials 18 02634 g003
Figure 4. The appearance of the electrolytic products (a), XRD patterns (b) and SEM–EDS results (c) of the cathodic products.
Figure 4. The appearance of the electrolytic products (a), XRD patterns (b) and SEM–EDS results (c) of the cathodic products.
Materials 18 02634 g004aMaterials 18 02634 g004b
Table 1. Weight loss rate and electrical resistivity of the furnace burden before and after carbothermal reduction.
Table 1. Weight loss rate and electrical resistivity of the furnace burden before and after carbothermal reduction.
Weight Loss Rate/%Electrical Resistivity/Ω cm
Before reduction/0.0440
After reduction26.910.0169
Table 2. Chemical compositions of Zr metal after washing.
Table 2. Chemical compositions of Zr metal after washing.
Batch NumberZr/wt.%Fe + Cr/wt.%H/wt.%O/wt.%C/wt.%N/wt.%
No. 199.50.180.0020.130.0210.012
No. 299.00.310.0040.190.0180.024
No. 399.10.270.0030.170.0240.019
Average99.2 ± 0.30.25 ± 0.070.003 ± 0.0010.16 ± 0.030.021 ± 0.0030.018 ± 0.06
R60704≥97.50.2 to 0.4≤0.005≤0.18≤0.05≤0.025
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

Song, W.; Chen, X.; Jia, Y.; Yang, M.; Ye, G.; Zhu, F. Preparation of Metallic Zr from ZrO2 via Carbothermal and Electrochemical Reduction in Molten Salts. Materials 2025, 18, 2634. https://doi.org/10.3390/ma18112634

AMA Style

Song W, Chen X, Jia Y, Yang M, Ye G, Zhu F. Preparation of Metallic Zr from ZrO2 via Carbothermal and Electrochemical Reduction in Molten Salts. Materials. 2025; 18(11):2634. https://doi.org/10.3390/ma18112634

Chicago/Turabian Style

Song, Wenchen, Xu Chen, Yanhong Jia, Mingshuai Yang, Guoan Ye, and Fuxing Zhu. 2025. "Preparation of Metallic Zr from ZrO2 via Carbothermal and Electrochemical Reduction in Molten Salts" Materials 18, no. 11: 2634. https://doi.org/10.3390/ma18112634

APA Style

Song, W., Chen, X., Jia, Y., Yang, M., Ye, G., & Zhu, F. (2025). Preparation of Metallic Zr from ZrO2 via Carbothermal and Electrochemical Reduction in Molten Salts. Materials, 18(11), 2634. https://doi.org/10.3390/ma18112634

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